Category Taking Science to the Moon

The Apollo Lunar Surface Experiments Package. and Associated Experiments

By 1964 the growing fraternity of space and lunar scientists began to see the Apollo missions as an opportunity to address many age-old questions. These questions related not only to the Moon itself but to the Earth, the entire solar system, and to some degree the whole universe. The Moon would provide the equivalent of a spacecraft on which to conduct experiments never before possi­ble. The Sonett Report, along with advisory panels from the Office of Space Science and Applications, the Office of Manned Space Flight, and the National Academy of Sciences’ Space Science Board, guided us in soliciting experiments to be associated with a permanent science station such as we studied for post – Apollo missions under contract at Marshall Space Flight Center (these studies became the basis for the Apollo Lunar Surface Experiments Package, or ALSEP, developed for Apollo missions). We also solicited additional experiments that could be conducted on the Moon’s surface independent of a geophysical ground station. At this time a few of the scientists who were thinking about experi­ments on the Moon were also considering how to conduct experiments in lunar orbit. Aside from Lunar Orbiter, however, there were no ‘‘approved’’ plans to provide a platform for lunar orbit experiments in the Apollo missions. I em­phasize ‘‘approved,’’ for though planning for such experiments was going on, no specific spacecraft had been designated to carry them. Experiments as well as cameras had been solicited for the Lunar Orbiter program, but the proposals were on the shelf until a program was approved.

Will Foster’s February 13, 1964, memorandum, in addition to recommend­ing a Field Geology Team that would help plan for sample collection, listed four geophysics teams, selected to recommend and design lunar seismic, magnetic, heat flow, and gravity experiments.1

The seismology experiment was divided into two parts, passive and active (each requiring different instrumentation), to monitor the Moon’s internal activity (moonquakes) and determine its shallow and deep structure. The team consisted of Frank Press, then at California Institute of Technology, and Mau­rice Ewing and George Sutton of Columbia University. The memo proposed additional investigators for the active experiment, but they were unnamed.

A third type of seismic experiment, engineering seismology, was also listed, to provide data that would be used for post-Apollo mission planning. Although considered a nonscientific experiment, it was designed to measure the Moon’s surface characteristics to a depth of fifty feet. For this team Foster suggested personnel from the Manned Spacecraft Center and the United States Geological Survey at Flagstaff, since USGS had begun to study the data needed for design­ing lunar bases and mobility devices under my office’s contract with Gene Shoemaker. The engineering seismology experiment was finally designated the active seismic experiment, and Robert Kovach at Stanford University became the principal investigator (PI), supported by coinvestigators Thomas Landers, also from Stanford, and Joel Watkins, who had moved from USGS at Flagstaff to the University of North Carolina. Kovach never selected anyone from MSC to join his team.

The magnetic measurements team consisted of James R. Balsley of Wesleyan University, Richard R. Doell from USGS, Norman Ness of NASA Goddard Space Flight Center, Chuck Sonett from NASA Ames Research Center, and Victor Vaquier from the University of California, San Diego. This team was to specify the magnetic measurements needed to determine the lunar magnetic field (if any) in the presence of solar and interplanetary magnetic fields and the methods for measuring any remnant magnetic field in lunar rocks. All previous attempts to measure a lunar magnetic field from a distance had failed to find any significant field; thus these measurements would be critical in unraveling the Moon’s early history.

The heat flow measurement team consisted of Francis Birch from Harvard, Sydney P. Clark from Yale, Arthur H. Lachenbruch of USGS, Mark Langseth from Columbia, and Richard Von Herzen from the University of California, San Diego. In addition to designing the heat flow instrumentation, the PI for this experiment would become closely involved with the design of the Apollo lunar drill, because the heat flow sensors would be lowered into two holes made by the drill.

The final team listed in the memo was to make gravity measurements. It consisted of Gordon MacDonald from the University of California, Los An­geles, and Joseph Weber from the University of Maryland. This experiment, it was hoped, would provide some of the more exotic measurements to be made on the Moon; not only would it measure the deformation of the Moon created by the pull of the Earth’s mass, but it might detect gravity waves, predicted by Einstein but never unequivocally measured. This experiment was truly unique to the Moon, since to have any hope of recording gravity waves the instrument, a sensitive gravimeter, had to be on an extremely quiet body, as many believed the Moon to be.

These teams, like the field geology, geochemistry, petrography, and miner­alogy teams, were also approved by the Space Science Steering Committee (SSSC). My purpose for listing the team members is twofold. First, it shows for the record that their members included many of the leading scientists of the day in the identified fields. Thus this obviated the need to make the usual formal solicitation to the scientific community as a whole, since it would undoubtedly have resulted in teams similar to those proposed. Some might take issue, but I believe that is true, since only a few leading scientists in these fields were considering lunar experiments. This procedure shortened the time it took to get the key players in place, probably by six months or more—not an insignificant consideration. Second, the makeup of the teams changed with time, especially the important position of PI for each experiment. This position, of course, was the key to future scientific fame, for the PI’s name would appear first in the final reports and citations.

Each team was to design and build its experiment through the prototype stage, training the astronauts in its use or deployment and, finally, reducing and reporting on the data returned. This opportunity was extremely important, because Apollo promised long-term data collection for experiments attached to the lunar ground station (one year or perhaps longer) and exciting data trans­mission (bandwidth) capability. Weight and power allowances were expected to be generous compared with a typical unmanned spacecraft experiment, and having an astronaut set up the experiment or return some or all of the data would add to the value. Some people in the unmanned science camp argued that using the astronauts for those types of experiments was an unnecessary complication, but in general their involvement was considered a real plus. Before this date in 1964 we had little experience deploying unmanned payloads either in space or on the Moon, and those that had been deployed were rela­tively unsophisticated. Using astronauts to set up or operate an experiment had only occasionally been factored into an experiment’s design for the Gemini program, so this would be a new challenge to the scientific community.

After SSSC approved Foster’s recommendations, contracts were negotiated with the team members, and OSSA began to fund and manage their efforts. As promised in Foster’s memo, other experiments and investigators were brought on board later to cover important areas of science not included in the initial teams. Experiments added during this time were the Solar Wind Spectrometer (SWS) to measure the solar plasma striking the lunar surface; the Suprathermal Ion Detector Experiment (SIDE), which could measure a variety of interactions with the solar wind and complement the SWS measurements; and the Cold Cathode Gauge (CCG) to measure the composition of the lunar atmosphere. These experiments would also be attached to the ground station for their power, housekeeping, and data transmission needs.

Another experiment that would operate independently of any ground sta­tion, the Solar Wind Composition experiment, was also approved for the first missions. It was proposed by a Swiss team headed by Johannes Geiss from the University of Bern. Its purpose was to collect and return solar wind ions, such as helium and neon, to help us understand the composition of the solar wind. This experiment was funded in part by the Swiss government.

With the initial suite of experiments and experimenters under contract, in early 1965 our efforts turned to the design and development of the station that would support the experiments. By this time the MSFC Emplaced Scientific Station (ESS) contractors, Bendix and Westinghouse, had progressed to a pre­liminary design of a geophysical station for post-Apollo missions. It had all the characteristics we wanted for an Apollo station; the major differences were in overall size, the ESS being larger than we could expect for the first Apollo landings.

On May 10, 1965, Foster sent Ernst Stuhlinger at MSFC a request to submit a work statement for an Apollo scientific station.2 At the same time he also asked Max Faget at MSC to submit a similar work statement. Much to our chagrin, George Mueller’s office, led by James Turnock and Leonard Reiffel, thought MSC should be the lead center in managing this complex payload. I was lobby­ing hard for MSFC and had convinced E. Z. Gray and Will Foster that, based on all the work MSFC had done for our post-Apollo mission studies, it was the best qualified. MSC had nowhere near as much experience with lunar science payloads, and it lacked qualified staff to oversee such a contract.

This controversy came to a head at a Saturday meeting with George Mueller, on May 24.3 (Remember the best day to get his undivided attention?) Also at the meeting (besides me) were Sam Phillips, Edgar Cortright (Mueller’s deputy), E. Z. Gray, Will Foster, Dick Allenby, Jim Turnock, Len Reiffel, Benjamin Milwitzky, and Jack Trombka. The major issue was deciding who would man­age the Apollo science station. We reviewed the two work statement proposals from MSC and MSFC and weighed the strengths and weaknesses of each. We described the problems we had working with MSC on matters dealing with science and the much better relationship we, and the scientists we had brought into the post-Apollo studies, had with MSFC.

After about two hours of discussing the pros and cons, the MSFC work statement was judged superior to MSC’s and likely to elicit the best proposals. I thought we had carried the day and that MSFC would be assigned this impor­tant task. But Phillips finally weighed in with his opinion—that in spite of its deficiencies MSC should become NASA’s ‘‘lunar expert.’’ Mueller agreed and also expressed his unwillingness to have Stuhlinger manage the Apollo science program. Why he was uncomfortable with Stuhlinger he never explained. He did agree that the MSFC work statement should be the basis of the request for proposals and asked that three companies be selected in the initial study to ensure some competition.

The anointing of MSC as our lunar expert was a devastating blow to many of us in attendance and presaged the at times bitter disagreements we and the PIs would have with the MSC managers in the years ahead. As a gesture to assuage our disappointment, Mueller asked us to prepare a history of our past year’s negotiations with MSC so that he could understand the situation. Perhaps Mueller’s review of our tales of past disagreements was a factor in the decision to transfer Bill Hess to MSC at the end of 1966 to lead the science activities there.

The MSFC’s Apollo work statement, based on the ESS study, in essence called for a junior ESS. Because extended periods of data collection were needed for many of the experiments selected, it was decided that the station would be powered by a radioisotope thermoelectric generator, the same power source proposed for the ESS. RTGs, already under development for planetary space­craft that would fly too far from the sun to collect sufficient solar energy, generate electric power through the decay of radioactive elements, in this case plutonium. This decay produces heat, which is in turn converted by thermo­couples to electric power. RTGs were an ideal power source for lunar-based experiments, because for fourteen consecutive Earth days of every day/night lunar month cycle, the station would be in darkness and very cold. Batteries alone could not do the job; they would be far too heavy to accommodate the required duty cycles. A solar-powered station would have required a large solar array, would be difficult to deploy on the lunar surface, and would still re­quire a large, heavy battery pack to sustain it during the fourteen-day lunar night. When the solar array and the batteries were studied, it became evident that RTGs provided not only a distinct savings in weight but also greater reliability and simplicity, because among their other advantages they have no moving parts.

The Atomic Energy Commission (AEC), the agency responsible for oversee­ing the manufacture of RTGs, had several well-tested designs to choose from that could provide various amounts of power depending on how much weight one could allocate to the power source. The RTGs were manufactured by Gen­eral Electric’s Valley Forge, Pennsylvania, plant, with the plutonium supplied by the AEC. Safety considerations were the primary arguments against using an RTG. First was the question of how an astronaut could safely deploy the RTG. It would be ‘‘hot,’’ both in temperature and in radioactivity. Second was the chance of a mission abort in which the plutonium 238 fuel source might be released into the atmosphere. Plutonium is highly toxic if inhaled. AEC and General Electric believed they could solve both problems, and later ground handling tests and destructive tests exposing the system to high-energy impact and heat loads proved them correct.

The RTG power source (system for nuclear auxiliary power, SNAP) selected to provide power for a year or longer was designated SNAP-27. It consisted of a fuel capsule and generator. The fuel capsule would be carried to the Moon on the lunar module descent stage in a special graphite container, and after the as­tronauts removed it and inserted it in the generator, it would provide 63.5 watts of electrical power to the central station. With a fuel half-life of almost ninety years, it more than filled the need for a long-term power source.

(The RTG on the Apollo 13 mission, still attached to the LM lifeboat that sustained the astronauts during that harrowing, nearly fatal experience, reen­tered at a speed far beyond that anticipated for a typical Earth orbit failure, but it is believed to have survived intact, as designed. If the cask protecting the plutonium heat element had failed, the sensitive instruments on the aircraft sent to sample the air at the reentry point over the Pacific Ocean would have detected plutonium in the atmosphere.)

Once the power source had been decided, the next critical step was selecting a design for the communication and data relay subsystem. Commands would be transmitted from Earth to control the station and its experiments, and data would be relayed back from the lunar surface. NASA’s Manned Space Flight Network (later incorporated into the Space Tracking and Data Network) would provide round-the-clock monitoring, eliminating the need to provide data storage at the station as we had envisioned for some of the ESS experiments. Raw or processed data would then be given to the PIs for reduction and inter­pretation. A difficult question was, How much data should the station be capa­ble of handling? No matter how much was made available, PIs would always be hungry for more. Until specific instruments were designed, this would remain an open question. At the Falmouth conference, attended by some of the proba­ble experiment PIs, it was recommended that the station be designed to accept various types of experiments so that the instrument complement could be easily changed, depending on the landing site and the experiments required to answer site-specific questions. All in all, it would be a tough design challenge, but based on the work we had done for our post-Apollo studies, we felt con­fident it could be met.

In June, using the MSFC work statement as its model, MSC released the request for proposal (RFP) for the geophysical ground station that came to be known as the Apollo Lunar Surface Experiments Package. Max Faget’s Engi­neering and Development Directorate’s new Experiments Program Office was named MSC program manager, reporting to Robert Piland, recently appointed to head the office. Nine companies submitted proposals, and as Mueller had requested, three were selected to provide a preliminary design. In August, Bendix Systems Division, Space General Corporation, and the TRW Systems Group were each awarded a contract for $500,000. Each company would pro­vide a preliminary design and mock-ups, to be delivered to MSC and Grum­man at the end of the six months.4

The RFP requested that each design include a seismometer, heat flow sen­sors, magnetometer, a suite of atmospheric and radiation sensors, and a device to measure the micrometeorite flux at the lunar surface. (This last device,

proposed by MSC and rejected by the Planetology Subcommittee four months earlier as not relevant to lunar science, had found its way back into consider­ation. MSC used its position as NASA’s ‘‘lunar expert’’ to push one of its pet ideas.) The weight allowance for the entire package was not to exceed 150 pounds. After a review and evaluation of each contractor’s design and perfor­mance during the six months, we planned to select one contractor to provide the flight hardware.

After the mock-ups were delivered, we convened a selection committee to decide which of the three teams would build the flight hardware. Bendix had obviously profited from its part in our post-Apollo studies of the ESS. Its pre­liminary ALSEP design was judged the most responsive to our requirements, and a contract was awarded in March 1966. With an initial value of $17 million, the contract finally grew to $58 million through increases in the number of flight and test units required and the added job of building four ALSEP experi­ments for the PIs and integrating more experiments than originally projected.

The contract and its subsequent modifications called for the manufacture of six flight-qualified ALSEPs, a ‘‘dummy’’ unit to fly to the Moon in the storage bay of the Apollo 10 lunar module, prototype and qualification units, two training units, and one unit dubbed the ‘‘shop queen,’’ which was modified and cannibalized and was generally available to test ideas. Joseph Clayton, later promoted to division general manager, was the initial Bendix program manager and was succeeded by Chuck Weatherred at the time ALSEP progressed to the prototype phase. Chuck, who had been closely involved with many of our post – Apollo studies, then continued as program manager through the missions.

Some additional details now about ALSEP, the attached experiments, and the other experiments that were deployed at the landing sites but were not dependent on ALSEP for their operation. First the ALSEP central station. The central station was the control center for the many instruments that were so carefully crafted by the experiment teams, some designed to record sensitivity levels unachievable for similar Earth measurements. The central station data subsystem would receive and decode the uplink commands for each experi­ment and collect the scientific and housekeeping data and transmit them back to Earth. A small helical S-band antenna would be mounted on top of the station and pointed by the astronauts to provide the data link to Earth.

Most of the experimenters were interested in collecting data over a long period, in most cases the longer the better. The ALSEP design goal was to survive for a minimum of one year, providing power, housekeeping functions, data collection, and transmission. This was no mean task, given the extreme temperature fluctuations (over 500°F) experienced on the Moon every twenty- eight days. At the same time that instruments or devices would be experiencing these temperature changes, they and the central station would be operating in a high vacuum. Lacking the normal methods of regulating experiment tempera­tures on Earth, their design would have to include novel ways to both heat and cool all the components.

Keeping the experiments warm was not as hard as keeping them cool; heat could be supplied by small electrical heating elements of various designs. But since liquid coolants could not be used, radiators, thermal blankets, and shield­ing were employed, utilizing new materials. In addition, the central station and the experiments would have to be carefully oriented to provide selective shad­owing and reflection of the sun’s rays.

Thirteen experiments were ultimately selected to operate with the five ALSEPs deployed on the Moon. (Some were on the ALSEP carried on the Apollo 13 mission, and their remains are at the bottom of the Pacific Ocean.) Each ALSEP had a unique combination of experiments, ranging from four to seven, and some experiments were carried several times. The eight listed at the begin­ning of this chapter were considered of highest priority. Four more would be added over the next few years, plus a dust detector to help monitor ALSEP’s health if dust or dirt on thermal surfaces caused a problem.

One of the four new experiments, Seismic Profiling, had an objective similar to the active seismic experiment but would provide additional information on the Moon’s shallow structure. The other three were the Lunar Ejecta and Mete­orites Experiment to measure the direction of travel, speed, and mass of mi­crometeorites arriving at the lunar surface (not the MSC proposal mentioned earlier); the Charged Particle Lunar Environment Experiment for measuring a wide range of charged particles caused by the interaction of the solar wind on the lunar surface; and the Lunar Atmosphere Composition Experiment, a spec­trometer that would measure the composition and density of whatever gas molecules might be found in the tenuous lunar atmosphere. Some of the experimenters did their own contracting and built their experiments, deliver­ing them to Bendix for integration, and some used Bendix as their contractor.

Nine other experiments, not dependent on ALSEP and not including those discussed in chapter 5, were to be deployed by the astronauts either in the

vicinity of the LM or during their traverses. They fell into two categories: those used for studying the Moon and those that used the Moon as a convenient platform from which to make measurements.

As I mentioned earlier, one could think of the Moon as a spacecraft circling the Sun, on which you could place instruments to measure phenomena occur­ring within or outside our solar system. In our post-Apollo studies we had examined using the Moon as a site for astronomical observations, and this preliminary study elicited some interest from the astronomy community dur­ing the Falmouth and Santa Cruz conferences.

On later missions, when larger payloads became available, we had the op­portunity to test this idea. An ultraviolet camera-spectrograph was proposed and carried on Apollo 16, the second J mission. The objective of the experiment was to evaluate the Moon as an astronomy base and to take pictures of targets in the far ultraviolet portion of the electromagnetic spectrum, a frequency that could not be studied from the Earth’s surface because of our intervening atmo­sphere. The experiment was proposed by George Carruthers of the Naval Re­search Laboratory, and the instrument was designed and fabricated at his lab.

A second experiment in the category of using the Moon as an observation post was the Cosmic Ray Experiment, a multipart experiment proposed by three teams, one at the General Electric Research and Development Center, a second at the University of California at Berkeley, and a third led by Caltech. Its objective was to detect high and low energy particles emanating from the Sun and from outside the solar system. It had the potential to record particles that had not been detected on Earth, again because of interactions with our protec­tive magnetic fields and atmosphere. It would go to the Moon mounted on the LM descent stage, where it would be exposed just after translunar injection, then folded and retrieved at the end of the third EVA. A related part of the experiment was a detector carried inside the astronauts’ helmets to determine their exposure to cosmic rays during their transit and stay on the Moon or while in lunar orbit. This was important information because it concerned the astronauts’ health, especially if a solar flare or some other major event that occurred somewhere in the universe at an earlier date would expose them to high energy particles during the mission. It was also important for planning longer-duration, manned missions to Mars.

Five of the nine experiments fell into the category of studying the Moon

through various measurements. These were the Lunar Neutron Probe, the Laser Ranging Retro-Reflector (LRRR), the Lunar Portable Magnetometer, the Lunar Traverse Gravimeter, and the Surface Electrical Properties (SEP) experiments. By the time the last three were proposed, it was known that a small vehicle would be available to the astronauts, so the magnetometer, gravimeter, and SEP were designed to be carried on the lunar roving vehicle (LRV), with measure­ments taken either at the astronauts’ discretion or at planned points. The magnetometer and gravimeter would measure the Moon’s gravity and magnetic fields to determine if these values changed as the astronauts moved away from the LM. The SEP used radio waves to penetrate the lunar surface to look for layering in the Moon’s crust. If there was no moisture in the upper layers, it might be able to penetrate deeper than the Seismic Profiling experiment. If water or ice occurred below the surface, the signals received would reveal their presence. The neutron probe would be lowered into the drill hole after the core was extracted to measure the rate of neutron capture below the lunar surface. This measurement would help us understand the physical processes that pro­duced the lunar soil. After remaining in the drill hole for some time, the probe would be recovered and brought back to Earth for study.

The LRRR was a late addition to the roster of Apollo experiments and deserves further description. A laser beam, originating at an observatory on Earth, would be reflected from the Moon back to the observatory and thus provide an accurate determination of the Earth-Moon distance (within a few inches). It was proposed by Carl Alley from the University of Maryland and was built in time to be carried on Apollo 11. Alley was supported by a host of coinvestigators; one of them, James Faller from Wesleyan University, became the PI for the Apollo 14 and Apollo 15 missions. The experiment was designed and developed by the A. D. Little Company of Cambridge, Massachusetts, and built by Bendix. The experiment carried on the Apollo 11 and Apollo 14 mis­sions consisted of one hundred reflectors, each about an inch and a half in diameter, arranged in a ten by ten square. They were mounted on an adjustable support that could be tilted and aimed at the appropriate angle for each landing site to best reflect the laser beams coming from Earth. The astronauts aimed the device using a sun compass and a bubble level, pointing the array at the center of the Earth. Individual corner-cube reflectors were manufactured under a separate contract by the Perkin Elmer Company. Because of difficulties in locating the LRRR at Tranquility Base and the Fra Mauro landing sites, the array carried on Apollo 15 was increased to three hundred reflectors, and it proved much easier to locate and reflect laser beams from Earth.

A network of three LRRRs was to be placed on the Moon, separated as far as possible in latitude and longitude. By sending laser beams from the Earth to the LRRRs and bouncing them back, it was anticipated that the Earth-Moon dis­tance could be calculated within approximately six centimeters. Such precise measurements would permit the study of many physical properties of the Earth and the Moon, including fluctuations in the Earth’s rotation rate, the wobble about its axis, the shape of the Moon’s orbit, and the Moon’s wobble about its axis. Ultimately, if enough stations on Earth were capable of sending laser beams to the Moon, small movements in the Earth’s crust might be measured. (Crustal movements are no longer measured this way. Instead, accurate laser ranging measurements are made from Earth to orbiting satellites. The LRRR, however, is the only experiment carried to the Moon by the Apollo astronauts that is still used for other types of measurements.)

Headquarters management of ALSEP was initially under the direction of OSSA in the Lunar and Planetary Programs office managed by Oran Nicks, and OSSA provided the funds to get ALSEP started. (The vast majority of ALSEP funding eventually came from OMSF.) William ‘‘O. B.’’ O’Bryant, a retired navy captain, was named program manager, and Dick Green, a retired air force officer returning to NASA after a recall, was named program engineer. Ed Davin, still reporting to Will Foster, was named program scientist. I also main­tained an oversight of ALSEP because of its close relationship to other programs I was managing, such as the lunar drill. Relations between headquarters staffers and MSC soured almost immediately. MSC continued its practice of not notify­ing headquarters when important reviews were to be held at Bendix or MSC. This caused a great deal of heartburn at headquarters. This attitude and way of doing business eventually led to the appointment of John ‘‘Jack’’ Small, who proved easier to work with; but the atmosphere had already deteriorated, and an uneasy relationship continued even when Small was replaced toward the end of the program by Donald Wiseman. Fortunately from our perspective, Bendix proved to be a cooperative contractor and recognized the importance of main­taining good relations with headquarters. This was a wise move, for in the ensuing six years there were a number of times when obstacles and difficult decisions arose that required the intervention of headquarters.

In some small defense of MSC’s reluctance to keep headquarters apprised of ALSEP’s progress, a careful line was always drawn between NASA’s contract offices and its contractors in order to avoid any misunderstanding about who was in charge. MSC had the sole authority to control the ALSEP contractor’s actions, and any changes to the contract scope could occur only with written direction from the MSC program manager. Probably all NASA centers had experienced instances when a contractor had used a conversation with someone from NASA headquarters to attempt to modify the scope of its contract, a surefire way to screw up the contract and make the center in charge see red. O. B., with his navy background, was a no-nonsense manager, and never to my knowledge did he create this kind of problem. But he never backed down from exercising his management prerogatives, which included the right to suggest changes to the program manager if he or his staff saw trouble developing and to keep a tight rein on the funding. O. B., Green, and Davin also felt that they were often the only ones sticking up for the experimenters when trade-offs were required, and they didn’t hesitate to make their concerns known.

Toward the end of summer 1968, with ALSEP development in its final stages, NASA management began to reevaluate the first landing mission’s lunar surface activities. Concern was growing about how well the astronauts would function on the Moon and, more important, how the LM would perform. Several ways to alleviate these concerns were explored. First the number and length of EVAs could be reduced. But if only one EVA was allowed, then ALSEP could not be deployed and still leave time for the astronauts to carry out their other important tasks, including sample collection. Not carrying ALSEP would reduce the astronauts’ workload and the weight of the LM for the first mis – sion—a partial solution to both concerns. Removing weight would also add a few seconds of hover time. ALSEP became a prime target for removal.

When rumors spread that the scientific experiments on the first landing would be drastically reduced, Charles Townes, chairman of the Science and Technology Advisory Committee, went to NASA senior management and ar­gued for keeping as much science as possible on the first mission. Our office, Bill Hess at MSC, and others in the scientific community were also lobbying hard to keep ALSEP on the first landing mission and to maintain two EVAs. Our office was fighting for more than just the science. ALSEP and the geological tasks the astronauts were scheduled to carry out represented years of planning and hard work, not to mention suffering through many a contentious meeting

with those in NASA who did not embrace the need to include science on Apollo. We were not prepared to accept such a defeat.

In September and October, in response to this outcry, our office and MSC studied an alternative to dropping the full ALSEP and presented it to the NASA Senior Management Council. A new, much smaller, and more easily deployed science payload was proposed and approved and given the name Early Apollo Experiments Scientific Package (EASEP). EASEP would comprise just three experiments, the passive seismometer, packaged with the dust detector, and the LRRR. Another self-contained, easily deployed experiment, Solar Wind Com­position, along with the equipment for the field geology experiment, would constitute the rest of the science payload. EASEP would be much lighter than ALSEP and require less time to deploy. By including these experiments on the first mission, NASA hoped to divert the criticism that was sure to come its way and show that its heart was in the right place regarding science. The astronauts would leave the highest priority experiment, the seismometer, at the landing site and still have time to conduct a limited geological study, collecting fewer samples than originally planned.

Instead of being powered by an RTG, the EASEP seismometer would get its power from solar panels and batteries charged by the solar array, the power source rejected for ALSEP but now acceptable because of the short lifetime expected for this substitute. The seismometer would have to operate only through the rest of the lunar day in which it was deployed, although we hoped it might survive longer. It would contain several small isotope heaters to help it survive the lunar night and, with luck, continue to function during a second lunar day. Like ALSEP, it would also have a self-contained telemetry system to transmit to Earth the seismometer and dust detector readings.

EASEP’s design was developed through close cooperation between MSC and Bendix, working under the ground rule that as much as possible of the hard­ware and subsystems would be based on ALSEP. Donald Gerke led the MSC team in the design phase and became the program manager for this hurry-up ALSEP substitute. In November 1968 NASA and Bendix agreed to a $3.7 mil­lion contract for the design and manufacture of the EASEP as well as the LRRR. By this time, with the Apollo flight program rapidly moving ahead, the date of the first landing was becoming obvious—sometime in the summer of 1969. Only three more Apollo test flights were scheduled before the first landing attempt. EASEP would have to be built in five months if it was to meet a May 1 delivery date at Kennedy Space Center for a subsequent June or July flight date. In contrast to some of the difficulties we encountered with MSC’s ALSEP managers, Gerke was easy to work with, especially for us at NASA headquarters. EASEP proceeded without a hitch and was delivered to KSC one day ahead of schedule.5

At the beginning of the chapter I listed the prominent scientists who were identified in Foster’s memo, along with the highest priority ALSEP experi­ments. In the months after the SSSC approved their selection to develop the experiments for Apollo, and before the ALSEP contract was signed, some ma­neuvering took place—at times a little indelicate—to determine who would be named principal investigators. This title conferred an important imprimatur because the PI would be the primary contact in the years ahead as we built the instrument and also would be responsible for interpreting the returned data and publishing the results. As a reward for all this effort, the PI would receive the largest amount of NASA funding allocated to the experiment and in some cases would be in charge of distributing funds to other members of the team. Remember the golden rule: ‘‘He who has the gold rules.’’ This was certainly the case for the PIs. In addition to the gold, they also got the publicity and all the other notoriety that went with this high-profile position. Most prominent sci­entists are not shrinking violets; being identified as Apollo PIs enhanced their reputations, and the exposure certainly helped advance their careers. How many scientists could look forward to saying they had designed an experiment that was placed on the Moon by the astronauts? Everyone knew only a lucky handful would have this claim to fame.

An example of the competition for this honor was the naming of the PI for the passive seismic experiment. Frank Press, while at Caltech, had developed the first lunar seismometer (which never flew) for Ranger. Maurice Ewing and George Sutton, at Columbia University’s Lamont-Doherty Laboratory, had de­veloped a seismometer (which likewise never flew) for Surveyor, and it was this very experience, plus their overall reputations, that led to their inclusion on the passive seismometer team. Ed Davin recalls a meeting at NASA headquarters to select the passive seismometer PI. Press and Ewing were present along with one of Ewing’s graduate students, Gary Latham. Ewing, being the senior scientist present, led the discussion and declared that Latham should be the PI be­cause this role would require that someone devote full time to the job and he thought—taking the liberty to speak for himself and Press—that they would not be able to do this. He volunteered that he, Press, and Sutton should remain as coinvestigators. Press, having studied under Ewing at Columbia University, graciously acquiesced, but after the meeting he remarked to Ed, ‘‘What Papa Doc wants, Papa Doc gets.’’ He was obviously disappointed at not being named PI by ‘‘Papa Doc,’’ a somewhat affectionate name given Ewing by his former students. Soon after, several others would be added to the team, but with Latham at Lamont-Doherty, Ewing’s laboratory reaped the public acclaim. Latham went on to do an outstanding job as PI.

I have not had the opportunity to talk to Press about this incident, but I imagine that in hindsight he might think it was not a bad decision. Soon after that meeting he became chairman of the Department of Earth and Planetary Sciences at MIT, certainly a full-time job. His reputation certainly did not suffer from not being an Apollo PI, for among other important jobs he held in later years, he was named president of the National Academy of Sciences, one of the most prestigious scientific positions in the nation.

There are some other interesting anecdotes concerning the experiments. Perhaps the most star-crossed was the Lunar Surface Gravimeter. Its tale of woe has been partially told before, but it deserves further discussion, perhaps with some new insights. I met the PI, Joseph Weber, early in his struggle to get his experiment accepted by NASA. His laboratory was only twenty minutes from our office in downtown Washington, on the campus of the University of Mary­land. Dick Allenby, Ed Davin, and I visited Weber in his basement laboratory sometime in early 1964. He had been building and modifying gravimeters in his lab for several years, hoping to arrive at a design sensitive enough to detect gravity waves. Gravity waves were predicted by Einstein’s general theory of relativity, and it was believed they could be generated by the collapse of some distant star or perhaps might emanate from an ancient supernova. It was believed that gravity waves would propagate outward from such events at the speed of light and that if one had a sensitive enough gravimeter they could be detected on Earth. Further, it was believed that analyzing them would provide new insights into the structure and evolution of the universe.

A secondary objective of his proposed experiment was to measure the defor­mation of the Moon by the tidal pull of the Earth. Weber showed us his latest model, and it was indeed a sensitive instrument—so sensitive that it was detect­ing large trucks and trains passing in the distance. Therein lay the snag that led

him to propose his experiment for an Apollo mission. The Earth was subject to so many events that would disturb its gravity field that some thought it would never be possible to make the delicate gravity measurements he wanted. The Moon offered a location without a lot of extraneous gravity sources—certainly no trains and trucks would mask gravity waves. Simultaneous measurements by similar instruments on the Earth and the Moon might identify movements that would be associated only with a passing gravity wave.

Weber’s experiment was eventually approved for Apollo, and he was given a contract to build a new gravimeter with the highest sensitivity possible based on the technology of the day (nominal sensitivity one part in 1011 of lunar gravity). He in turn contracted with Bendix to build his instrument with a subcontract to LaCoste and Romberg, world-famous builders of gravimeters, to design and supply the sensor. Because of the late approval to include the experiment on Apollo 17 and the complexity of the design, MSC questioned in August 1970 whether the experiment could achieve delivery in July 1972. We suggested shortcutting some of the normal procedures and, if necessary, integrating the flight hardware with ALSEP at KSC instead of Bendix.6 Development pro­ceeded on this new schedule with just the usual problems one encountered in such a program, and the flight instrument was delivered on time for integration with the Apollo 17 ALSEP, the last opportunity to get it to the Moon on an Apollo flight. Because its objective was so unusual, it was billed as the star experiment of the Apollo 17 mission. Weber and his coinvestigators, John J. Giganti, J. V. Larson, and Jean Paul Richard, all from the University of Mary­land, eagerly anticipated being the first to unequivocally detect the elusive gravity waves. Gordon MacDonald, originally on the team with Weber, had dropped out, for reasons I don’t recall.

Astronauts Eugene Cernan and Jack Schmitt, aware of its scientific signifi­cance, practiced diligently with the training model to be sure they would not foul up its deployment. In his pamphlet On the Moon with Apollo 17, printed just before the mission, Gene Simmons, MSC’s chief scientist, went so far as to predict that ‘‘the practical utilization of gravitational waves may lead to benefits that far exceed those gained from the practical utilization of electromagnetic waves’’ (italics in the original). That would be a hard prediction to live up to, but his pronouncement reflected the enthusiasm and anticipation that accom­panied the gravimeter to the Moon. An article in Science in August 1972 reported that a race was on at a number of laboratories around the world to be the first to confirm the measurement of gravity waves, labeled an ‘‘exotic problem.’’7

On the Apollo 17 mission ALSEP and the gravimeter were deployed on the first EVA by Jack Schmitt, approximately six hundred feet west-northwest of the LM. When commands were sent to the gravimeter to turn on the experiment, readings were received almost immediately back in the Science Support Room. The readings didn’t look right to those monitoring ALSEP, but this was the first time the instrument had operated in the reduced gravity of the Moon, so no one was quite sure how the signal should look. Jack completed the ALSEP deployment and activation and went about his other tasks. Meanwhile the Bendix engineers and Weber tried to figure out how to get the instrument to operate more to their liking. They tried to rebalance the beam (the part of the sensor that responded to the pull of gravity) by sending commands to add or subtract mass on the beam, but the signal coming back didn’t change signifi­cantly with these commands.

A ‘‘tiger team’’ was appointed to come up with a solution while the astro­nauts were still on the Moon and might be able to help resolve the difficulty, although at the time it still wasn’t clear what the problem was or how serious it might be. Perhaps just a little rap by one of the astronauts might clear up what appeared to be a minor discrepancy in the instrument’s readings. Schmitt went back to the gravimeter several times during later EVAs to jiggle it a little, but the instrument still did not respond as expected. The beam seemed to be resting on the upper stop and not swinging free. Jack’s comments reflected his concern that perhaps he had made some mistake during the deployment, but he had done nothing wrong.

When the Apollo 17 astronauts left the Moon, Weber and the Bendix engi­neers were still unhappy with the gravimeter’s readings but could not find the cause. Perhaps operating the instrument through one or more lunar day/night cycles might help clear up the signal; so it was monitored for the next several months, but there was no change in the response. The Preliminary Science Report for Apollo 17 came out almost a year later still promising that the gravimeter would return useful information. But it wasn’t to be.

Back at Bendix, in Ann Arbor, a second team delved into the mystery. The instrument had been checked out at Bendix before shipment and had worked satisfactorily. What had gone wrong? Then it occurred to LaCoste what had happened. To test the gravimeter on Earth a set of weights were dropped on the balance beam, correctly calculated for Earth’s gravity. After the test these weights were recalculated to compensate for the Moon’s gravity, which is much less than the Earth’s (1/81), and installed by LaCoste. Because of a faulty calculation, those installed were not the proper weights for the Moon.8 Thus this experiment, on which so much hope for a major discovery had been riding, never returned much useful data. Joe Weber and his team of coinvestigators never forgave LaCoste for the mistake. Perhaps accelerating the schedule con­tributed to this miscalculation, but at the time it seemed a reasonable risk to get the instrument on the final mission.

At this time, to my knowledge, no one has directly detected gravity waves, and new efforts are under way to build a gravity wave experiment called LIGO.9 LIGO’s announced objective is to detect gravity waves originating from black holes or supernova events. Sound familiar? Designed by scientists at several universities and funded by the National Science Foundation, two identical instruments have been built. One has been installed at the Department of Energy’s Hanford, Washington, laboratory and another at Livingston, Loui­siana. The two instruments will permit simultaneous measurements at distant points, thus removing the possibility that, rather than signaling the passage of a gravity wave, the mirrors used to bounce a laser beam back and forth in a tunnel two and a half miles long would be misaligned by some local disturbance (such as the trucks and trains observed in Weber’s earlier experiments). A difficult quest, but perhaps this time it will succeed.

Another experiment that caused a problem was the lunar surface magne­tometer (not to be confused with the Lunar Portable Magnetometer). In this case Chuck Sonett, the PI, chose to have the instrument built by Philco-Ford and then integrated by Bendix. (The PI on the Apollo 15 and Apollo 16 missions was Palmer Dyal, also from the NASA Ames Research Center.) The sensor electronics for the instrument contained thirteen hundred active components, eighteen hundred passive components, and thirty-three hundred memory core locations. It included thousands of tiny diodes supplied by Fairchild. Scheduled to fly on the first ALSEP, with the first landing fast approaching, all the ALSEP experiments were under pressure to meet the schedules for delivery, test, and integration at Bendix. Prototype instruments were always tested before build­ing the actual flight hardware to ensure that the design would perform as advertised, and the tests were closely monitored by MSC and headquarters.

When the prototype magnetometer was tested it failed miserably. Short circuits were noted at many places in the circuitry. Trouble. Was there a major flaw in the experiment design? And if so, would there be time to redesign to meet the schedule and have a new instrument ready for this high priority experiment?

The prototype was torn down and subjected to a careful analysis that re­vealed the problem. To meet the tight schedule, the circuits had not been properly cured, or ‘‘burned in,’’ and in addition many of the diodes were contaminated by fine particulate matter embedded in the potting compound. Fixing the curing time was easily solved, but how did the contamination occur?

A team from headquarters and MSC went to Fairchild to review its man­ufacturing techniques, and the contamination mystery was solved. After the diodes were manufactured, they were placed on shelves—not in a clean room— to cure. Dust and other airborne contaminants circulated in the air and stuck to the potting compound, and these minute particles were enough to permit arcing across the circuits. But could Fairchild come up with a new batch of clean diodes in time to meet the schedule? With the first ALSEP deployment postponed until Apollo 12, Fairchild pulled out all the stops and made the delivery, saving the magnetometer’s assignment. The instrument operated suc­cessfully for many months, with only a few minor discrepancies that were corrected as it continued sending information back to Earth.

Five years after the last Apollo mission, at the end of September 1977, Noel Hinners, who had left Bellcomm and later had been appointed NASA associate administrator, Office of Space Science, decided to save the $1 million per year spent monitoring the five ALSEPs and sending the data to the PIs. Few data were being recorded by this time. It was not expected that the passive seismic experiment, probably the most interesting experiment still operating, would provide much new information because there were no more man-made im­pacts on the horizon, and naturally occurring major meteor impacts and large moonquakes were uncommon.

During the years they were operating, before being put in a standby condi­tion, all the ALSEPs were still functioning long past their design goals, though occasional glitches and data dropouts were observed. Before NASA terminated support for the ALSEPs, several engineering tests were conducted on the central stations and some of the experiments. These tests were devised to answer questions raised during their operational lives but that had not been allowed to be asked for fear of damaging the ALSEPs and the experiments. The test results were then filed away for possible use if another ALSEP-like station was sent to the Moon. After these tests, commands were sent to the ALSEPs, with each of the PIs sorrowfully taking part in the ceremony, to place their experiments and central stations in a standby mode in case someone wanted to turn them back on later. In the meantime, no data would be collected or transmitted.

In October 1994 the Department of Energy (the successor to the AEC, which had provided the RTGs) wanted to determine if the RTGs had survived over the intervening eighteen years. Ground controllers at the Johnson Space Center tried to reactivate the stations. They hoped the ALSEPs would still be receiving power, as predicted by the plutonium half-life, waiting to spring back to action when Earth called. They made two attempts to turn on each of the ALSEPs, but none of them responded. Although the RTGs were probably still generating electric power, it seems likely that as the RTGs aged and power levels dropped, the ALSEPs turned themselves off, as designed, when a minimum operating power level was reached.10 The next time they are revisited will probably be when some intrepid lunar explorer or entrepreneur lands near an Apollo land­ing site and drives over to recover pieces, as we did for Surveyor 3 during the Apollo 12 mission, bringing them back to be put in a museum or someone’s private collection of space memorabilia.

Walk, Fly, or Drive?

Safety was always the primary concern when someone recommended the astro­nauts carry out an action. As new ideas were suggested, the astronauts were included as early as possible so they could offer their point of view. When the debates began on how to provide mobility on the lunar surface, they made their thoughts known decisively. The best lunar surface transportation mode would have to take into account not only their preferences but also the payload weight available on the lunar module, the tasks to be performed, and the equipment the vehicle would have to carry. Those looking through the narrow lens of the Field Geology Team wanted the astronauts to cover as much ground as possible at each landing site and carry a variety of tools for mapping and sample collec­tion. The geophysicists and other science disciplines, as we saw at the Falmouth and Santa Cruz conferences, had their own particular requirements for deploy­ing experiments and collecting data. For the Astronaut Safety Office, the pri­mary concern would be to keep the astronauts always within easy reach of the LM in case any of a wide variety of emergencies occurred.

An astronaut walking on the Moon would be, in effect, a small, self- contained spacecraft. His space suit and all the attached systems would have to let him function in the brutal lunar environment (high vacuum, low gravity, and extreme temperatures). It could be as cold as —260°F in shadow, while in full sunlight a short distance away it might be 270°F. He also had to see objects and the ground around him both in shadow and in the glare of the full sun. While moving about he would need a way to maintain voice communication with Earth and, ideally, automatically relay information on his physical condi­tion and the status of his life-support systems so those monitoring them could tell him if he had to return to the LM. Designing a space suit that would accommodate all these multiple functions was an enormous challenge for the Manned Spacecraft Center engineers and their contractors. My office and the scientific community followed their progress with great interest, for the more successfully these challenges were resolved the more scientifically productive the missions would be.1 The astronauts had to be mobile, and they had to maintain good eye-hand coordination; the closer space suit designs came to allowing “shirtsleeves” efficiency the better, though we knew that could not be achieved.2

The space suit solution for the Apollo missions was based on technology developed in the United States and Great Britain, first for pilots flying high – altitude fixed-wing aircraft and, more recently, for the Mercury and Gemini programs. The MSC Engineering and Development Directorate and the Crew Systems Division directed the efforts of many contractors, some retained from Gemini, to produce the Apollo extravehicular mobility unit (EMU), the com­bination of suit and attached support systems. Hamilton Standard and Inter­national Latex Corporation were chosen as the prime contractors for the EMU design and manufacture.

The major elements of the EMU were a liquid-cooled inner garment to re­move body heat; an eighteen-layer outer suit, topped by an integrated thermal – meteoroid cover lest a tiny meteorite punch a hole in the suit; a helmet with a clear inner visor and a sunshade (added after Apollo 14) and a movable, trans­parent gold-plated sun reflector visor; gloves; and boots. The portable life – support system (PLSS), attached to the back of the space suit, included bat­teries, fans, pumps, and the expendables (oxygen, water, and lithium hydroxide canisters to remove carbon dioxide) plus a separate oxygen purge system con­taining thirty to seventy-five minutes of oxygen in case of a failure in the PLSS.3 All together, the EMU weighed about 200 pounds (60 for the suit and 140 for the PLSS), varying with the mission and the additions or improvements it embodied. The EMU went through several upgrades from Apollo 11 to Apollo 17, each designed to improve the astronauts’ ability to perform their tasks on the lunar surface.

Perhaps most difficult to design were the gloves. I attended several design reviews over the years as improved glove designs, incorporating new materials, were demonstrated. At each review the technology improved, although some ideas were discarded as development proceeded. The gloves had to be tough enough to confine the suit’s internal gas pressure (3.7 psi) in the lunar vacuum and to withstand abrasion from handling rocks and equipment. At the same time, the gloves had to allow the astronauts some sense of touch. These two requirements worked against each other from a materials point of view: high wear strength and toughness resulted in poor feel through the gloves. Imagine trying to thread a needle wearing work gloves with the fingers blown up like balloons. Not an exact analogy, but pretty close.

The final design had an outer shell of tough fabric covered with thermal insulation and fingertips made of silicone rubber so the astronauts could feel what they were touching. Not a perfect solution, but the best the technology of the day would permit. In spite of the attention given to this part of the suit, the astronauts would often end their simulations, or return to the LM after a long stint of extravehicular activity on the Moon, with bloody fingertips, cracked fingernails, and their hands aching from trying to grasp and hold a wide variety of objects. However imperfect, the glove design did the job. No glove failures occurred during the missions, and all scheduled tasks were completed.

The EMU restricted how the astronauts could perform various tasks, how far they could wander from the LM, and how long they could stay outside the LM on any EVA. The suit and backpack mass would have to be large, the equivalent of moving a heavy weight with every step. In addition, the astronauts would be continuously working against the internal suit pressure to bend the suit at its joints. Walking on the Moon would thus be difficult and tiring despite the low lunar gravity. If an astronaut fell it was feared he might not be able to get up, and the difficulty was accentuated because the PLSS, attached at shoulder height, raised his center of gravity. (This proved not to be a problem; in the Moon’s low gravity, the astronauts could easily bounce up from a fall.) But EVA planning required that they always be close enough together to help each other if one should have a problem. The PLSS provided for sharing oxygen and cooling water if one PLSS malfunctioned.

While suit development was under way, these restrictions raised the specter that the astronauts might not accomplish the demanding work being planned during the lunar EVAs. Metabolic tests had been made on many suited test subjects as well as on several astronauts simulating the tasks to be done on the Moon.4 Data from these tests showed that the EMU then available would limit EVAs to four hours of low level work. The PLSS could supply consumables (the oxygen, water, and lithium hydroxide mentioned above) for four hours if the astronauts averaged a metabolic rate of 1,200 BTU/hr, the equivalent of playing golf in shirtsleeves. If they exceeded this rate they would have to reduce their activity to reach the average use of consumables if the EVA was to last the full four hours. In reality this would mean almost standing still, since just moving slowly in the suit required over 1,000 BTU/hr; 600 BTU/hr was needed just to work against the suit’s internal pressure and overcome joint friction. In spite of improvements in the Apollo EMU during the next few years, the results of these analyses led, in part, to a decision to reduce the amount of EVA time on the first landing mission. EMU consumables were carefully monitored on all missions, especially when the astronauts undertook tasks not programmed in the mission timelines.

These considerations also led to continual upgrades of the Apollo suit and research on better space suits. In May 1968 Sam Phillips asked MSC to recom­mend a program for space suit development with an eye to improving the astronauts’ mobility on the lunar surface for the post-Apollo missions. (He wanted the improved suit to be ready by 1971.) An EVA working group, report­ing to Charles W. Mathews, Mueller’s deputy associate administrator, began meeting to look into all aspects of EVA, both in free space and on the lunar surface.5 Ames Research Center became involved, since it also had a team working on space suits; its favorite was the constant volume suit, a hard suit like a deep-sea diver’s suit. James Correale led the work at MSC’s Crew Systems Division and coordinated the MSC research with that going on at Ames. Many of the concepts combined properties of the soft and hard suits, including articulated bearings, bellows joints, and metal fabrics. Although it promised to reduce the astronauts’ workload, the hard suit never was adopted because of operational considerations, including the extra stowage space required. How­ever, the hard suit, or a hybrid suit, is still under consideration for Space Station EVAs because it reduces metabolic demands. Perhaps when materials science improves and spacecraft design permits its use, it will be adopted as the stan­dard EVA suit.

For Apollo 15, Apollo 16, and Apollo 17 several suit improvements were made, including making it easier to bend at the waist and adding expendables (water, oxygen, lithium hydroxide, and a larger battery) to the PLSS to allow longer EVA time-all important improvements for these missions. Since EVAs for these missions might last as long as eight hours, the pressure suits also provided a few creature comforts, with an emphasis on ‘‘few.’’ Most important for such long EVAs, bags containing one quart of drinking water were attached to the helmet neck ring inside the suit. The astronaut could reach a straw by turning his head inside the helmet. A small snack bar also could be attached to the neck ring and eaten by turning the head.

At the other end of the human system, a urine bag was attached inside the pressure suit leg to collect urine, much like the earlier “motorman’s friend’’ for trolley car operators. Back in the LM the urine bags would be removed from the suits, and later they would be left on the Moon. Now you know the answer to one of the questions people most often asked the astronauts. The other adjust­ment made for the final three missions was that some of the tools could be attached to the pressure suit or PLSS so the astronauts did not have to return to the lunar roving vehicle (LRV) to retrieve them from the tool carrier during their sample collecting and geological studies.

EMUs used on the lunar surface EVAs differed from those worn by the command module pilots; beginning with Apollo 15, they had to make an EVA to retrieve film and tapes from the experiments bay of the service module during the return trip from the Moon. The CM pilot’s EMU did not include the PLSS; it was attached to the CM by an umbilical cord that supplied life-support consumables and voice communication links. The EMU did include a small emergency backpack containing the oxygen purge system, similar to that at­tached to the lunar surface EMU.

With the Apollo suit being developed, studies described in chapter 3 were already under way at Marshall Space Flight Center on two alternative types of vehicles: flying machines and motorized wheeled vehicles. The wheeled vehicles were championed by most members of the science community, led by the Field Geology Team at Flagstaff, and were supported by my office at NASA headquar­ters, while the flying machines were favored by some of the staff at MSC and a few astronauts. Our simulations at Flagstaff had used many types of wheeled vehicles, and procedures and operations that took advantage of a vehicle were far advanced. Based on this work, the choice seemed obvious; the astronauts should be equipped with some sort of wheeled vehicle.

Lunar flying vehicle (LFV) proponents at MSC were basing their support on the work that Textron-Bell Aerospace Company had completed at MSFC, also described in chapter 3. The LFV engendered visions of astronauts zooming above the lunar surface like Buck Rogers, free to go wherever they wanted, and quickly. Clearly the LFV would be able to reach places a wheeled vehicle could not go. But would the astronauts be permitted to use such a device, considering safety concerns and the possible need to walk back to the LM from dangerous locations if the LFV failed? Discussions during the Falmouth conference were not supportive of it as an exploration tool. Mission simulations using a flying vehicle were never carried out in the field owing to the difficulty and expense of providing a good simulation. Only Textron-Bell pilots were qualified to use the LFVs, so based on a few demonstrations by the manufacturer, one had to imagine how such a vehicle could be used on the Moon.

This debate came to a head at the Santa Cruz summer conference in August 1967, with heated discussions between the two factions. As is often the case in government matters, when opposing positions are strongly held there are no clear winners, and this was true at Santa Cruz. The final report endorsed both wheeled vehicles and flight concepts. Since we were focusing on post-Apollo missions (in 1967, planning for the first Apollo landing missions envisioned only the astronauts’ walking), we were not constrained from advocating robust vehicles, going so far as to recommend using both types to jointly support the surface exploration. In spite of this accommodation at Santa Cruz, momentum was building in favor of a wheeled vehicle for the later Apollo flights. The recommendations coming out of the several working groups called for contin­uous traverses, manned and unmanned, to sample and deploy various types of equipment and experiments, operations that did not lend themselves to a flying machine.

In April 1969 Frank Press, who had chaired both the Falmouth and Santa Cruz geophysics working groups and was now a member of the Lunar and Plan­etary Mission Board (LPMB), submitted a paper representing the board’s lean­ings and recommending a ‘‘lunar exploration program.’’6 Only three months short of the first lunar landing and still anticipating ten lunar landings, Press’s paper emphasized the need for enhancing mobility: first, with a better space suit to improve the astronauts’ walking and overall EVA capabilities, and second, with some type of wheeled vehicle operating in both manned and unmanned modes to ‘‘interpolate between type locations.’’ In Press’s words, with increased mobility, the strategy outlined in the paper ‘‘provides optimal scientific return and fully exploits the Apollo capability.” The LPMB unanimously approved this recommendation at its next meeting in May and passed it on to Homer Newell.

With concerns about the astronauts’ ability to move about on the Moon plaguing Office of Manned Space Flight management, George Mueller stepped in and made a decision. The argument of ‘‘fliers’’ versus wheeled vehicles was finally put to rest, and the wheeled vehicle won. Safety was probably the critical factor in the decision. If a lunar ‘‘jeep’’ broke down, the worst result would be a long walk back to the LM. If a flying vehicle had a problem it might crash in an inaccessible area. Other considerations were also important, such as stowage and the overall weight of a fully fueled flier (more than three times as heavy as a projected lunar ‘‘jeep’’) that could carry two astronauts many miles. As envi­sioned by the Santa Cruz attendees, the LFV would complement a surface vehicle; but as a stand-alone or only means of transportation, the LFV was too limited to support the planned science, especially for the final missions, when multiple EVAs were planned that would include many geophysical measure­ments at many points along the traverses. Because the LMs had limited payload capacity, a choice had to be made, and the LRV won.

Mueller convened his Senior Management Council in May 1969. At the meeting, attended by George Low, at that time MSC’s Apollo spacecraft pro­gram manager, and Wernher von Braun as well as other senior OMSF man­agers, Mueller asked Low and von Braun to examine the problem and arrive at a solution. A small LRV was the final choice, and Mueller told Sam Phillips to go ahead with it. At the end of May Phillips sent a memo to MSFC, the center with the most experience in lunar vehicle research, asking it to manage the procure­ment. Von Braun wanted an experienced senior manager to lead the effort, and he tapped Saverio ‘‘Sonny’’ Morea to be the program manager. Morea had not been in on any of the earlier MSFC lunar roving vehicle studies, but he had been program manager for the Saturn У F-1 engine development, a critical and difficult job that he had successfully completed. He had been given a ‘‘heads – up’’ for his new assignment and had attended the Senior Management Council meeting.7

With Morea’s appointment, the procurement was put on a fast track. Ben Milwitzky, who had just finished his role as headquarters’ manager of the Surveyor program, was transferred to our office to oversee this new program. Ben was a good choice because at the beginning of the Surveyor program a small wheeled vehicle was a candidate payload (though never flown), and Ben had several companies under contract working on their concepts. He had some hands-on experience to guide him in developing the larger vehicle for the Apollo missions.

In July MSFC released the request for proposal (RFP), and three companies responded—Bendix, Grumman, and a Boeing-General Motors team. We all thought Bendix had the inside track to win the contract because of its involve­ment in all the post-Apollo vehicle studies, plus it was the only one of the three bidders that had a working model of its concept at the time the RFP was released. Boeing also had a good background because of its work in post-Apollo studies, having teamed with General Motors (Delco Electronics Division) for the mobile laboratory competition. Grumman believed it would have an ad­vantage because it had done some earlier work on a one-man vehicle. The design of this new vehicle would be intimately tied to the LM and its stowage constraints, and of course no one knew the LM better than Grumman.

After the Source Selection Board (SSB) reviewed the proposals, it deter­mined that Bendix and Boeing had the superior proposals and passed its find­ings to NASA headquarters. Because of the short schedule-seventeen months from projected contract start to delivery of the flight vehicle-headquarters told MSFC to negotiate contracts with both companies, not knowing which one would be chosen by the source selection official, Thomas O. Paine, the new NASA administrator. With negotiated contracts in hand, we would be able to jump-start the contract and save valuable time. Of the two bids, Boeing had submitted the lower price, $19.7 million, and since all the other SSB findings were essentially equal, Paine awarded the contract to the Boeing team.

MSFC then signed a performance-based contract (a wise decision, as it turned out) that went into effect in November 1969. Included on the Boeing- GM team were Eagle-Pitcher Industries, which supplied the LRV batteries, and United Shoe Machinery Corporation, which provided the electric harmonic drive units that powered each individual wheel. It would be a true four-wheel – drive vehicle. The contract called for the delivery of four vehicles (later reduced to three) and six test units, one of which was eventually converted into a one-g trainer for astronaut simulations on Earth.

Soon after the contract went into effect, MSFC and headquarters had some misgivings about the specifications contained in the contract. Morea’s team thought they were too complex and opened the door for possible change orders that would boost the price and perhaps jeopardize the schedule. For example, the original RFP called for a gyroscopically controlled navigation system. After careful review, the high accuracy this type of system would deliver was thought to be unnecessary, and it would add to the overall cost. On January 15, 1970, Ben chaired a meeting of engineers from MSFC, MSC, and Kennedy Space Center to rectify this situation and develop a less restrictive set of specifications.

The design requirements coming out of that meeting, and then translated into the final specifications for the Boeing team, called for an LRV that would carry one or two astronauts plus experiments, communications, a TV camera, and crew equipment and would provide stowage for lunar samples collected during the traverses—a total payload capacity of 970 pounds.8 In place of the gyroscopic navigation system, it would have a rudimentary system that would give the astronauts a continuous vector back to the LM in case it was out of sight and they needed to make a rapid return. Other specifications called for the LRV to travel a maximum of ten miles an hour on level mare surfaces with an overall range of seventy-two miles.

The most demanding requirements were that the vehicle be transported to the Moon in the wedge-shaped LM descent stage Quadrant I and that the total weight of the vehicle, including its stowage and deployment mechanisms, could not exceed four hundred pounds. This meant the LRV would have to be folded or collapsed and that the chassis and wheels would be flimsy indeed.

After all the vehicle studies we had performed for the post-Apollo missions, I was skeptical that the overall specifications could be met within the weight and stowage constraints. This would be smaller and lighter than anything we had studied for post-Apollo, yet it was being designed to accomplish many of the jobs we had envisioned for our larger vehicles. I shared my concerns with Ben, but he was convinced the specifications were valid. Events proved that such a vehicle could be built with these tight constraints. I credit his management skills, along with the dedication and engineering know-how of Sonny Morea’s team plus the hard work and cooperation of Boeing, GM, and their suppliers, for the on-time delivery of the LRVs—the payload stars of the last three Apollo missions.

The LRV team encountered many complications as it struggled to meet the tight schedule. Early in the contract, MSFC concluded that the Boeing program manager did not have the skills to manage such a critical program and asked that he be replaced. Boeing agreed and brought in a new manager, Edward House, who took control and saw the project through to its successful conclu­sion. The next problem was the escalating cost. Congress got wind of this and asked the Government Accounting Office to review the contract. Here the performance-based contract proved valuable, because MSFC could demon­strate that the contractor’s rising costs were justified, based on the LRV’s design complexity, and that the contractor fee (profits) would be adjusted accordingly to arrive at the best price for the government. At a hearing at which Milwitzky and Rocco Petrone, who had recently replaced Sam Phillips as Apollo program director, testified, they explained the way the contract worked. They were able to satisfy the House Oversight Committee that the costs were realistic for such an unusual vehicle. The matter was dropped, and the final cost, with modifica­tions to the original contract for the LRV flight and test units, was just under $37 million—a bargain in the opinion of all who were involved in the missions.

While the LRV was in development, two new data points were thrust into the discussions on astronaut mobility. The first was the comments of the Apollo 11 astronauts after their return. Although their EVAs had been reduced in number and length so that their total time on the surface was just a little over two hours and thirty minutes, Neil Armstrong and Buzz Aldrin came back with the im­pression that walking on the Moon would be easy. They had discovered that a loping, rolling gait was the most efficient way to move and helped overcome some of the space suits’ deficiencies—in particular the difficulty of bending at the joints. Armstrong said he thought an LRV would not be needed to get around and to conduct the tasks the scientists had planned. When Morea asked at one of the debriefings what size wheels he would recommend to ensure that the LRV could handle surface irregularities, Armstrong replied, ‘‘about twenty feet.’’9 His opinions carried some weight, but in the end they did not slow the development of the LRV, and a much smaller wheel (sixty-four inches), did the job.

The second, more positive data point was the experience of the Apollo 14 astronauts. For Apollo 14 we had built a small two-wheeled cart called the modularized equipment transporter (MET) that the astronauts would pull along loaded with whatever equipment they needed during their traverses and that would also store the collected samples. By this time the array of geological tools and sampling devices we wanted the astronauts to carry had grown con­siderably, including three cameras. As Alan Shepard and Edgar Mitchell strug­gled to reach the rim of Cone Crater, the primary sampling objective of the mission, the MET became a bigger and bigger hindrance. In the end, as they tried to climb the slope to the crater rim pulling the MET behind them, they decided it was easier to carry it. Walking and pulling even a small cart created such a high workload that the astronauts often had to stop and rest before continuing their exploration. Because of the extra effort expended attempting to reach the rim, and with time running out, they were forced to return to the

LM, and they never quite reached their objective, though they came close. There seemed to be no question that with the much more ambitious missions next on the schedule, we were right to insist on having a motorized vehicle to carry the astronauts and their equipment.

By the time the first LRV was delivered to KSC on March 15, 1971, two weeks ahead of schedule, some of the original specifications had changed. Overall weight had been allowed to grow to 460 pounds, and its allowable payload had also grown, to 1,080 pounds. Its total range had decreased from seventy-two miles to forty. The reduction in range was acceptable as new mission rules developed for the LRV traverses dictated that the astronauts stay within six miles of the LM so they could walk back if the LRV failed.

Television pictures and voice communication would be possible from the LRV at the limits of the traverses, out of sight of the LM. A self-contained lunar communications relay unit would be carried on the LRV or could be hand carried. The LCRU would provide a direct link to Houston by two antennas mounted on the front of the LRV. The low gain antenna would permit voice relay with only coarse pointing toward Earth, but the high gain antenna, re­quired for TV transmission, had to be pointed rather accurately by the astro­nauts. This meant that voice communication would probably be available throughout an EVA, but TV pictures normally could be transmitted only when the LRV was stopped or when driving if the antenna happened to be pointing toward Earth. The LCRU would also permit a operator at Mission Control to point and focus the TV camera when the astronauts were working away from the LRV. The first LRV would be available starting with Apollo 15, and we were waiting with great anticipation for the TV pictures from the new LCRU. It promised the flexibility to monitor and communicate with the astronauts that we had tested in our post-Apollo simulations at Flagstaff.

Edward Fendell, who got the nickname ‘‘Captain Video,’’ trained for many hours to operate the TV camera from his station in the Mission Operations Control Room during our Apollo simulations and had become adept at manip­ulating it to get the best coverage. This skill was invaluable to the ‘‘back­room’’ Field Geology Team, and Ed cooperated to the fullest with their re­quests for views of the local topography at each stop. The media, especially the TV networks, were also excited about closely observing the astronauts at work and broadcasting live the promised spectacular scenery of the last three landing sites.

As a bonus, the LCRU would let us witness an LM takeoff from the Moon. At the end of the last EVA, the astronauts would drive the LRV about three hun­dred feet from the LM and park it with the LCRU on board and the TV camera pointed toward the LM. If Fendell could coordinate elevating the camera with the liftoff, we would be able to watch the LM disappear into the black lunar sky. Despite the difficulty of slewing the camera fast enough to follow the rapidly accelerating LM, Fendell accomplished this feat. At the end of the Apollo 15 mission, the world saw for the first time a slightly blurry view of a spacecraft taking off from another body in our solar system. We were also able to see the effects the LM’s ascent engine exhaust plume had on the lunar surface and the Apollo Lunar Surface Experiments Package. It was a little frightening for the ALSEP engineers to see debris flying in all directions, but the ALSEP survived. If the LCRU still had enough battery power after the Apollo 15 astronauts left, we hoped to take pictures of the lunar eclipse that would occur a week later (assuming the launch stayed on schedule, which it did), as well as other views of the lunar surface and astronomical targets. These observations were success­fully carried out.

A few final words will describe the LRVs, the remarkable machines that made Apollo 15, Apollo 16, and Apollo 17 so successful. The wheels were con­structed of an open wire mesh, to reduce weight, make it easy to stow in the small LM bay (the wire mesh was compressible), and damp the ride by flexing and acting as shock absorbers as the LRV bounced across the lunar surface in the low gravity. The open mesh had some drawbacks, however; as was correctly predicted, the wheels picked up soil and sprayed it over the LRV and the astronauts, so each wheel was covered by a small fender to direct the spray downward. (On Apollo 17 one of the fenders came loose during the first EVA traverse, and the soil spray coated the LRV and the astronauts’ space suits and equipment with a thick layer of dust. The next day Gene Cernan and Jack Schmitt made a new fender by taping together stiff sheets from their landing site maps and attached them over the wheel. Even so, when riding on the LRV or just walking around, the astronauts would return covered with lunar soil that they had to brush off before reentering the LM.

The LRV’s front and back wheels could be steered together, in tandem, or each pair independently, allowing it to make tight turns. It was steered with a small T-shaped hand-grip controller, which also regulated speed and braking. A knob below the T-handle controlled forward and reverse, much as in a golf cart.

Mounted above and just forward of the T-handle was the control and display panel, which contained a speedometer, LRV system switches (e. g., for power and steering), temperature gauges, and the onboard navigation system. This last system provided a continuous bearing and range back to the LM and also showed the total distance traveled to help the astronauts find their predeter­mined science stops.

All in all, the LRV was a dandy little machine that performed flawlessly. Full – scale models can be seen at several NASA centers as well as at the Smithsonian Air and Space Museum, which also displays a lunar module mock-up and other examples of equipment the astronauts used. If—or when—we go back to the Moon, it would surprise me if small vehicles similar in appearance and per­formance to the Apollo LRV are not part of the equipment included in the payloads. Why pay to redesign such a successful system? I hope Boeing or NASA has kept the drawings.

Astronaut Training and Mission Simulation

Just before I arrived at NASA, in April 1963 the United States Geological Sur­vey had reached an agreement with the Manned Spacecraft Center to start a geological training program for the astronauts. Ellington Air Force Base, a few miles west of the proposed location for the main MSC campus and home of the NASA astronaut air force, was selected as the site for this rump USGS office. Gene Shoemaker chose Dale Jackson, a former marine, to lead this effort, thinking his background would allow him to mesh successfully with the astro­nauts, who were all military pilots. Until that time the astronauts were not perceived as enthusiastic about studying geology, in view of their other pressing duties. By the time I joined NASA, stories were already circulating that some MSC staff members and Jackson’s small team did not agree on who was to call the shots on this important function. MSC staffers believed they should be in charge, although USGS had been given this mandate by NASA headquarters. Adding to the problem, the newly hired MSC staffers assigned to work with Jackson’s people did not have as much experience as Jackson’s staff, yet he agreed to include them in the training. As in other areas I have described, MSC had a pronounced fear of being left out of important assignments related to Apollo science and tried whenever possible to monopolize these roles.

In spite of the friction between the two staffs, Jackson plowed ahead with his duties and devised classroom and fieldwork courses in basic geologic princi­ples, mineralogy, and petrology. With the astronaut office’s approval, the syl­labus called for fifty-eight hours of classroom lectures and four field trips. The fifty-eight hours of ‘‘geology’’ training were part of an overall classroom syl­labus of 239 hours designed to prepare the astronauts for the upcoming Gemini flights.1 The geology training was not related to the upcoming Gemini flights, the astronauts’ primary concern at that time, and would not have real value unless they were selected as Apollo crewmen. Thus it was not universally em­braced, especially by some of the original seven and the second and third astronaut classes. Eventually, however, it became accepted as an essential box to be checked off if one hoped to be chosen for a Moon mission. It was anticipated that after crews were selected for the lunar landing missions, five additional series of follow-on lectures and field trips would be scheduled.

By 1967, one hundred hours of classroom lectures and ten field trips became the requirement for astronaut geology training. This training, and then the mission simulations, would become more and more rigorous and realistic as the program matured and simulations were scheduled using prototype and final design equipment and tools.

Three weeks after joining NASA in September 1963, I attended my first demonstration of a prototype Apollo space suit at MSC. The demonstration and briefing were done under the auspices of MSC’s Crew Systems Division. Hamilton Standard had been awarded the overall contract to develop the Apollo space suit and backpack, with International Latex, its subcontractor, responsible for the suit itself. This was my first opportunity to see the current state of the art in space suits. The prototype Apollo suit we were to see demon­strated was the latest amalgamation of this technology, plus modifications added by the Crew Systems staff, which had the ability (or expertise) to second – guess the contractor and make its own adaptations when appropriate. At this point two types of suits were under consideration: a ‘‘soft suit’’ made of multi­ple layers of nylon and other material and a ‘‘hard suit’’ to be made of some type of hard plastic or honeycombed aluminum material. This was a ‘‘soft suit’’ demonstration, the preferred approach.

A test engineer wearing the suit went through a series of mobility exercises for the assembled throng. Some movements he could carry out easily; others were more difficult or almost impossible. Bob Fudali and Noel Hinners of Bellcomm also attended the demonstration and filed a detailed report on what they had observed. They wrote: ‘‘All in all, it looks as if mobility will be rather low (even in improved suits) and that the astronauts will not travel far from the LEM without additional mechanical aids. [Their] ability to set up equipment and perform experiments on the surface will also be quite limited unless strik­ing changes are made in future suits.’’2 I also reported in a memo to my office what I had seen and what I believed were the deficiencies in the design.

My first exposure to astronaut training and simulation came at the end of August 1964 with a trip to Bend, Oregon. At this early date many had ques­tioned the astronauts’ ability to carry out meaningful scientific observations and work on the lunar surface while encumbered by the available space suits. I was one of the skeptics, based on the earlier space suit demonstration at MSC. My report on the 1963 demonstration had gotten back to Max Faget’s office at MSC and was considered so negative that when MSC found out I would be attending the Bend simulation, Faget sent a telegram to Tom Evans disinviting me. Ed Andrews told me to ignore the telegram and go anyway.

The Bend simulation, supported by several MSC offices, was designed around a space-suited astronaut, Walt Cunningham, alternating with two MSC technicians in space suits. They would work at several locations, using a few rudimentary field tools, and at the same time report what they were doing and seeing. The Bend location was chosen because it seemed like a good terrestrial analogue of what the astronauts would find on the Moon. It consisted of three types of volcanic terrain. One site was primarily a field of basaltic extrusives, jagged and rough and in places containing pieces of obsidian. MSC, it was rumored, was considering using the area as a permanent simulation site. Gover­nor Mark O. Hatfield (not yet a senator) and the press had been invited to witness parts of the simulation, and the exercise rapidly turned into a major public relations gaffe.

During the simulations, Walt wore the prototype Apollo space suit demon­strated less than a year earlier, with a few improvements including a new back­pack. It was the best suit available at the time. Together the suit and backpack and a bulky white overgarment weighed more than a hundred pounds. It was a blazing hot day, uncomfortable even for those of us just standing and watching in shirtsleeves. Walt’s suit was fitted out with a new water-cooled inner garment, best described as a pair of long johns with a network of thin plastic tubes sewn on. Cold water circulating through the tubes was supposed to keep him from overheating. It didn’t. His visor often fogged over, and he had trouble seeing where he was going.

One slope he tried to climb was covered with pieces of razor-sharp obsidian, and as might be expected, he tripped and sliced a hole in one of his gloves. Before this he had tried to use a geologic hammer and scoop to pick up samples. Both tasks were awkward in such a garment, but to make matters worse he had to carry the tools in one hand or hung at his waist and at the same time manipulate either a “walker” or a ‘‘Jacobs staff’’ that was supposed to help him conquer this rough terrain. At every stop he would put down the walker or staff and begin his next task. No matter how hard he tried, every action looked difficult. Whenever he bent over he tended to lose his balance because the suit was not designed to bend easily at the waist, a deficiency we had noted a year earlier. After he fell and cut his glove he continued to tumble down the slope and was saved from injury only by two technicians standing nearby just in case. All in all, it was a simulation disaster, which the local press reported the next day in large headlines.

By the end of the simulation, with a short rest after his fall while the tear in his glove was repaired (‘‘duck tape’’ helped get us to the Moon), Walt attributed his problems to his fogged-over visor and other suit limitations. He described the scene to his superiors back in Houston as a ‘‘Roman holiday,’’ referring to the swarming photographers eagerly taking pictures of his pratfall. Bob Fudali of Bellcomm also was there to observe the simulation. In his report he noted that ‘‘predicting the mobility of an astronaut on the lunar surface from these tests would be a serious error.’’3 My report to my office also retold Cunning­ham’s mishaps, and when copies of our memos were brought to his attention, he came to associate us with his bad press. The main points of our memos had been to argue for a suit that would make the astronauts more mobile and for better-designed tools, not to criticize Walt’s efforts. This simulation was an important factor that led him to caution us at the Falmouth summer con­ference not to overload the astronauts with lunar surface science tasks. Later I was able to explain my position to him and we became good working partners, though Walt never quite forgot his embarrassing Oregon experience.

My report also addressed the disadvantage of having such a large public attendance at simulations where many new things would be tried for the first time. I recommended that future simulations be done at Flagstaff, where we were beginning to set up good facilities and where attendance might be con­trolled. I had, of course, an additional motivation: to legitimize the role USGS was playing in our post-Apollo simulations and put the staff in a position to more strongly influence what would be done for Apollo. Will Foster and E. Z. Gray agreed with my suggestion, and each sent a memo to George Mueller recommending that Flagstaff be the future site for simulations.4 The Office of Space Medicine also sided with our observations and recommended policies to guide future simulations, including that astronauts ‘‘not be used as test sub­jects’’ unless they would make some unique contribution.5 Mueller forwarded these memos to MSC. He got back a letter from George Low, deputy director at MSC, disagreeing with Foster and Gray on their recommendation to conduct future field simulations requiring special terrain at Flagstaff and claiming there was no intent to set up a ‘‘lunar training camp’’ at Bend.6 This last statement played down Governor Hatfield’s comments while he was at the simulation that he supported having such a ‘‘camp’’ at Bend. This seemed to confirm the rumors we had heard that MSC had indeed made some preliminary overtures. It was clear that Low was telling Mueller they intended to do their own thing, especially when dealing with USGS.

Low’s response prompted Foster to send Mueller another memo to clear the air; he said that his earlier memo was not intended as a criticism of MSC but repeated his concern that pressure was being exerted on NASA to establish a training facility at Bend.7 To put an end to this internal bickering, Mueller wrote to Bob Gilruth, the MSC center director, ‘‘It is my desire that the Centers work closely with the USGS. . . and that there be no unnecessary duplication of field simulation activities,” and he sent an identical letter to Wernher von Braun at Marshall Space Flight Center.8 This exchange, unfortunately, only deepened the growing animosity between MSC and our headquarters-USGS team.

As field geology training picked up speed and our post-Apollo studies pro­gressed, we were constantly trying to find sites that would demonstrate terrains similar to those we expected the astronauts to encounter on the Moon. USGS already had a selection of sites it used at different stages in the training program, depending on the objective. Training trips took the astronauts to many distant places, both in the United States and overseas. But as our understanding of the Moon grew from pictures returned by Ranger, Surveyor, and Lunar Orbiter, new sites that could mimic the lunar surface were in demand for both Apollo and post-Apollo mission planning.

In May 1964 Bill Henderson, Don Elston, William Fischer of USGS, and I went hunting for sites that might be suitable for simulating longer missions and lunar base activities. Final reports from Bill Henderson’s Lunar Exploration Systems for Apollo (LESA) lunar base studies were due in nine months. Interim reports were already suggesting a broad range of undertakings that could be carried out at a base, and we used these early reports as a starting point for planning lunar base simulations. In those heady days we were thinking big; a lunar base program would undoubtedly be announced in the near future, to follow the successful Apollo missions. Until this time simulations for post – Apollo missions had been conducted exclusively near Flagstaff. We were look­ing for one or more large sites, not too remote and preferably on government property, where we could expect to find support for the lunar base simulations, which we anticipated would be complex. We drew up a list of potential loca­tions, obtained photographs and other background material, and reduced the large number of candidates to a short list.

We went first to the Atomic Energy Commission’s Nevada Test Site (NTS), where a series of surface and subsurface atomic and high energy chemical explosive tests had pockmarked the landscape with craters of all sizes. The local AEC manager was interested in our proposal, and though the site had restricted access, some sections could be made available for training. We were given a helicopter overflight, and from the air there was no question that it appeared moonlike. One crater, called Sedan, was especially impressive. Formed by a 104 kiloton explosive, the crater was 320 feet deep and 1,280 feet across. Flying over it at low altitude reminded me of standing on the rim of Meteor Crater in Arizona, for it had many of the same characteristics. After we landed we toured the site by truck to get a closer view. When we got out of the truck at the first stop, we discovered a major problem; we had to put on white coveralls and boots because the surface soil was still slightly radioactive; the atomic clocks of some of the products of the nuclear explosions were still ticking. We should have expected this situation, but when we made our calls to set up the tour, the fact was not mentioned. We looked at each other and rolled our eyes, then after a few short excursions we thanked our hosts politely and left.

Our second stop was China Lake, a large navy test range in southern Califor­nia. We studied a large-scale map of the range at the headquarters building and selected a few spots for a close-up truck survey. The range was vast (1.1 million acres), with lots of room for the many exercises we were hoping to conduct. Although it was not as Moonlike as NTS, vegetation was sparse and there were many interesting geological formations that could simulate lunar conditions. We toured the range by truck and agreed that it looked like a good site, and the commanding officer seemed willing to accommodate us. The test range also included many shops, hangars, and other facilities that we would need to support long-staytime simulations. They could be made available, we were told, with appropriate compensation.

From China Lake we next visited Fort Huachuca, Arizona. After a meeting with the commanding general, who assured us of his interest, the army also provided a helicopter overflight, followed by a series of briefings on facilities and other advantages of working there. They were definitely selling: perhaps they saw reduced budgets in their future and thought this new use might offset these reductions. This army proving ground was beyond question isolated. The Huachuca Mountains formed the western border of the fort, and a variety of volcanic terrains could be found within its boundaries. Although the region was semiarid, it was a ‘‘green desert.’’ Most of the ground was covered with cactus, including cholla, palo verde, and other types of plant life common to the area; it was beautiful, but we thought it would be too difficult to cope with continuously for sustained long-distance walking and vehicular simulations.

Our final stop was the White Sands Missile Range in south-central New Mexico. It was similar in many respects to China Lake. There was lots of space, some areas had Moonlike terrain, and there were good support facilities. NASA was already using some of the range, so we would not be unwelcome guests. It was perhaps the best of the sites we visited. As events unfolded, we never had to make a choice. Lunar base funding and planning came to an end about a year later, and our more modest post-Apollo simulations were all carried out near Flagstaff.

We continued to look for additional Apollo training sites, however, and a new tool became available to assist us. On each Gemini flight the astronauts took photographs of the Earth’s surface with handheld Hasselblad cameras. Many showed areas never before well documented with aerial photographs. For each flight Paul Lowman, with his coinvestigator Herbert Tiedemann at MSC, had designated points of special interest that the crew should try to photo­graph, time permitting. Gemini missions were launched due east from Kennedy Space Center to take full advantage of the extra boost from the Earth’s rotation; thus their flight paths repeatedly covered all of the Earth’s surface from 28.5° north latitude to 28.5° south. One of the benefits of repeating the launch inclination was that it was possible to rephotograph the designated areas when the photos from earlier missions were of poor quality or were not taken. This also allowed some stereoscopic coverage where the photos overlapped.

Using these photos, Paul and I searched for other potential training sites. Each Gemini photo typically covered an area of some 3,500 square miles, with the oblique photos covering even more—an unprecedented continuous view of the Earth’s surface. In the typical aerial survey, an average frame might cover less than ten square miles. Conventional photographic coverage of the large areas included in a typical Gemini frame would require constructing photo mosaics, with trained photogrammetrists piecing together many separate pho­tographs. Having used such products in our geological pasts, we knew that no matter how skillfully fabricated, photo mosaics always introduced false infor­mation in the finished maps. A geologist could be misled by something that looked like a stream or valley or some geological feature such as a fault but was really an edge between two photos.

Features never fully photographed before the Gemini missions, such as the Richat structure in Mauritania, that might be the result of large meteorite impacts were of special interest because they might provide not only training sites but also the opportunity to learn more about impact processes. In 1965 only a few well-documented impact craters were known throughout the world, and many of them were so obscured by erosion that they were not well suited as training sites. Thus we were constantly trying to find more examples that we could study or use to train the astronauts.

A few of the Gemini photos had been published in National Geographic, Life, newspapers, and other publications, but the vast majority had not been seen by the general public. In his spare time Paul had been carefully cataloging the pictures and interpreting their geologic content. It occurred to us that these new views of the Earth might interest companies exploring remote parts of the world. So far, no commercial interest had been shown. If we could get a positive response, it would support NASA’s proposed Earth orbital remote sensing program-just in an early planning stage—and perhaps persuade NASA man­agement to accelerate this program.

In May 1966 I called Mobil Oil in New York and talked to my old boss, James Roberts, who had been transferred after I left Colombia, first to Venezuela and then to Mobil headquarters. I explained what we had and what we thought would be the potential benefits and applications of space photography. He said he was interested in seeing the photographs and agreed to set up a meeting with some of the Mobil Exploration staff, the unit responsible for finding new oil fields. A few weeks later Paul and I flew to New York to show the Gemini photos to their first commercial audience. We brought to the briefing some of the best examples of geological features photographed by the astronauts; mountain ranges in the southern Sahara (Mobil was heavily involved in exploring remote areas in Libya and Tunisia) and clear pictures of structures in Iran of the type petroleum geologists looked for (anticlines and synclines). I knew Mobil had several field parties working in Iran at that time, because before I left Colombia Iran was a possible new destination for me. We also included a few spectacular views of the Andes and the Himalayas. We felt sure there were no aerial photo­graphs of some of these areas, and this would be the first time Mobil had such views available. We thought they would be impressed.

We were wrong. For whatever reasons, the staff members Roberts brought to our meeting showed little interest. They said they had, or could get, enough conventional coverage so that space photographs were not needed. This re­sponse mystified us. Perhaps they thought an endorsement would leave them open to providing financial support for an undertaking with an uncertain future. We will never know what might have happened if Mobil had been enthusiastic. Like other programs that were struggling to get started at this time, the Earth orbital observation program limped along, in part because there was no strong commercial interest. It would be many years before the unmanned Landsat program and Skylab would be launched.

Our search for terrestrial impact structures took us on two trips, one back to Colombia in April 1964 and another to Peru in June 1968. We visited Colombia to study a small circular structure of unknown origin, Lake Guatavita, high in the eastern cordillera of the Andes, some thirty miles north of Bogota. Lake Guatavita was an intriguing and well-known feature; at the time of the Spanish conquest it was rumored that the Chibcha Indians, who lived on the high plateau that surrounds what is now Bogota, used the lake for special cere­monies. It was said that the local chief would cover himself in gold dust every year and then bathe in the water, accompanied by other sacrificial ceremonies. The Spanish had dredged the lake and attempted to drain it in hopes of finding sunken treasure. A modern attempt, again unsuccessful, had also been made to drain the lake after several marvelously intricate gold artifacts were recovered from the bottom. Geological study had failed to come up with a satisfactory explanation of the lake’s almost perfectly circular shape; one suggestion was that it was created by an impact, but no proof had been reported. I had visited the lake while living in Colombia and was aware of its history and the impact theory.

Now that there was better understanding of how to identify an impact crater in the field, Paul and I developed a field study plan for making a quick assess­ment of the lake and submitted it for approval. The estimated cost of the trip for the two of us, including all expenses, was $1,000. In the memos that went back and forth before approval was given, a number of interesting comments were appended to the routing slips. The most humorous was one made by George Mueller’s special assistant, Paul Cotton: ‘‘George, this is the slickest justification for a boondoggle I have ever seen. As long as we have this kind of resourcefulness, we should be confident of reaching the moon and planets.’’ A second staff comment to Mueller was that approval should be given only if we included an astronaut. We were in favor of this recommendation, but it was soon shot down as taking too much valuable astronaut time. Our ‘‘resourceful­ness’’ was rewarded, and the trip was approved.

Our plan was to quickly survey the lake’s immediate surroundings looking for evidence of impact in the form of shatter cones or other impact debris such as ejecta, glass, or meteorite material. For two days we tramped around the half­mile-diameter lake picking up samples, taking pictures, and making a few measurements. We could find no evidence of an impact. This left us in a quandary: How should we report our results when there was so little to report? We felt sure that thin-section study of our samples would only confirm our field observations that the lake was not the result of impact. We went back to my old Mobil office in Bogota to examine more closely what was known, geologically, of the immediate area. Based on the published literature, we concluded that since we could find no evidence of an impact the lake was probably formed when the surface rock collapsed over a small salt dome that had been dissolved by groundwater. Thick salt deposits were known to exist in the underlying formations, and a complete cathedral had been carved below ground from the salt at Zipaquira, a short distance away. And so we reported our findings.9

When E. Z. Gray forwarded our report to Mueller we received a short handwritten acknowledgment: ‘‘I doubt if the returns were worth the time and money. Do you agree?’’ Gray wrote back: ‘‘What value do you place on develop­ing an organization? I am a firm believer in learning by doing. I think this trip was worthwhile.” Although it was only a small incident in a rapidly accelerating major national undertaking, this story provides a measure of the attention to detail demonstrated by senior management and at the same time the freedom of action they allowed their staffs. Such management competence, and such security in their abilities, may have had no equal in a government program before or after and was, I believe, instrumental in Apollo’s success.

The Peru trip was instigated by our study of the photographs returned by

Gemini 9. During the flight the astronauts had photographed the Andes from Chile to Colombia. At the point where the mountain chain turns from a mostly north-south direction to the northwest near Lake Titicaca in southern Peru, we observed several large circular structures, each having a diameter of thirty miles or more. Were they created by impacts or by some other mechanism?

After Paul and I found the circular structures on the Gemini photographs, we tried to determine if they had been discussed in the geological literature. We found no citations. Such large structures, if formed by impacts, would be a major discovery. We could see many large impact craters on the Moon, and by this time we had in hand the detailed Lunar Orbiter photographs that showed some of the fine structure associated with large impacts. We knew of no impact craters of this size on Earth, although we were sure that, like those on the Moon, they had been made during the planet’s early history. The Ries Kessel structure in Germany, about fifteen miles in diameter, which was used as an astronaut training site, was the largest confirmed terrestrial impact feature known at that time. The Vredefort Dome in South Africa, some twenty-five miles across, was potentially a larger example but was yet to be studied in detail. Many aspects of the large lunar craters were intriguing, especially their central peaks. Only large lunar craters had such peaks. Why did they exist? Did they reflect the thickness of the lunar crust or some other unknown phenomenon? The Gemini photos showed that the large circular structures in Peru had mountains in their cen­ters. We started to lay plans to visit Peru and try to answer our questions on the origin of these features.

As our planning progressed, Paul could see it would be difficult for him to make the trip; he had returned to Goddard Space Flight Center and new duties. I continued to pursue the idea and finally received permission to go from my new boss, Lee Scherer. In preparation I had been in contact with the United States and Peruvian embassies as well as the Peruvian Geological Survey and was assured of their cooperation. From the Defense Intelligence Agency I had obtained aerial photographs of the area taken in 1955 so I could plot our findings in the field. Interestingly, these relatively high resolution individual photographs gave no indication of the structures, and a photomosaic made from these photos also failed to show them. The advantage of the small-scale space photos, which covered a large area without distortion, was clear. In addition to these rather formal arrangements, I received an unexpected bonus. A NASA colleague, Rollin Gillespie, who worked in the Planetary Missions

Office, was interested in joining me. His son Alan, who was majoring in geology at Stanford, was also interested; so Rollin, at his own expense, offered to meet me in Lima and accompany me along with several Stanford students.

I arrived in Lima on June 15 sans baggage and field equipment, lost some­where en route. Rollin and his group had arrived several days before and had been in touch with the Peruvian Geological Survey. He had already made arrangements for two Land Rovers and for drivers, guides, translators (Spanish to Quechua), and three Peruvian geologists to accompany us. This saved us several days, since I arrived on the weekend and could not have made such connections for two days. While waiting for my baggage we met with the minerals attache at the United States embassy and with several other organiza­tions that were conducting mining operations in the area, and they supplied important information about the conditions we would encounter. An engineer at the Madrigal Mining Company told us they were working several large copper and silver mines in the center and on the flanks of two of the structures. This was encouraging; perhaps these circular features were similar to the Sud­bury structure in Canada, thought by some to be the remains of an impact crater, which was being mined for nickel, copper, and other metals.

Our plan was that Rollin and I would fly to Cuzco, where we would be joined two days later by the rest of the party and the Land Rovers and then travel south to the site. We flew to Cuzco on schedule and met, as we had arranged, with geologists at the National University of San Antonio to explain our project. They had never seen the Gemini photos and were excited by them. They were familiar with the region but had never realized these circular structures existed. While visiting at the university we received our first bad news. The rest of the party had been delayed in leaving Lima and would not arrive for several days. We decided to have them bypass Cuzco and meet us at Sicuani, a town near the base of the mountains. Before leaving the university I promised to stop on my way back to Lima and lecture to faculty and students on the Apollo program.

The next day Rollin and I took a bus to Sicuani, the only ‘‘gringos’’ on a bus filled to capacity with local passengers and all their baggage, some of it alive. It was essentially a straight shot through the Vilcanota Valley, which connects Cuzco to the altiplano that surrounds Lake Titicaca. Sicuani lay some eighty – five miles south of Cuzco by way of unpaved roads but with some spectacular scenery along the way. We arrived in Sicuani late in the afternoon and checked into the only hotel (warm water available every morning from 7:00 to 7:30). It was very cold. Sicuani is at an elevation of 12,000 feet, and there was no heat in the rooms, where we spent an uncomfortable night. By chance, while walking in the main plaza that first night, we met an American Carmelite priest who invited us to the parish house, where we discussed our plans with the assembled fathers. We then received our second round of bad news. They had visited the general area and told us it was not possible to drive in—it was too rough and there were no roads. We would have to rent horses. This would certainly slow up our exploration and add more time than I had available. They suggested we enlist the bishop’s support.

We met Bishop Hayes the next morning, and he was very helpful. Not only did he understand local politics and know who could ease the way, but he had a large, comfortable house (hot water all day) where he invited us to stay. We immediately agreed. The rest of our party arrived the next day, and we com­pleted our arrangements for renting horses and obtaining other equipment. With the delays in getting started my time in Peru was running out. I would be unable to travel to the structures and would have to depend on Rollin and the Stanford students, along with the Peruvian geologists, to complete the survey.

Returning to Cuzco by train, I stopped for the afternoon to deliver a lecture at the university. From Cuzco I flew back to Lima and then home. Back at NASA, I received a package from Professor Carlos Kalafatovich V. on the staff at the university in Cuzco. It contained several Peruvian newspaper clippings noting that scientists from NASA had visited the region and were interested in the mountains near Sicuani. According to the papers, which featured big black headlines that translated to ‘‘Flying Saucers Land in Canchis’’ (a small town near Sicuani), some of the local people interviewed were intimately familiar with those mountains. It seems that the locals knew of frequent visits by flying saucers that came to extract precious gems from somewhere in the mountains and take them back to their home planet. Now we knew what had attracted us to these structures.

On a more serious note, the party I left behind was not very successful. It was almost impossible to travel in the mountains, even using horses. They collected a few samples and took them back to Stanford for analysis. They found nothing unusual, and no sign of impact was observed in the mineralogy of the returned samples. The origin of the circular structures was not solved, and as far as I know the question is still open.

Backing up a bit, in September 1965 I participated in one of the astronaut training trips to Medicine Lake, California, a site near several small, complex volcanic features. By this time astronaut training trips were well organized by USGS and included prominent geologists who could lecture and teach the astronauts about the importance and subtleties of the locations selected and about their potential similarities to lunar features. This was the second two-day trip astronauts made to the area, and those on this particular trip were Russell ‘‘Rusty’’ Schweickart and Roger Chaffee. Roger was soon to be named to the crew selected to fly Apollo 1, scheduled to be the first manned flight of a Saturn rocket. Gene Cernan was also scheduled for this trip, but because of a hurricane threat he was delayed in Houston and unable to attend.

Roger Chaffee had come to the astronaut corps from the navy and held the rank of lieutenant commander. Since we were both jet pilots with many similar interests and experiences and had flown off some of the same class aircraft carriers, we hit it off immediately, and he became my truck mate for the training trip. I drove, and between scheduled stops and lectures I would fill him in on geological lore I thought he should know. But as I remember, we mostly swapped sea stories about night carrier landings and the idiosyncrasies of the planes we flew. He seemed to welcome the change of pace from his ‘‘normal’’ astronaut assignments, even though each day he was subjected to nonstop lectures and fieldwork while being force-fed textbook geology.

The team assembled for this trip consisted of ten people. Aaron Waters led the team and was to deliver the lectures and coordinate the trip itinerary. He was supported by nine helpers, including three USGS camp hands, two USGS geologists, two MSC geologists, and two MSC photographers. The astronauts’ doings were always well documented by photographs. Dick Allenby and I were also invited for this trip, so there were fourteen of us. We slept in one – or two – man tents and were up at dawn to complete each day’s tightly scheduled busi­ness. Breakfast was served around a campfire because the early morning hours were already chilly. At noon we had box lunches, and dinner was back at the campsite. This trip turned out to be especially memorable because William Rust, one of the USGS ‘‘camp hands’’ but in reality a technician, was the designated cook and an inveterate fisherman. Each morning, before any of us were awake, Bill would go to the lake and catch trout, then cook them for breakfast—a treat in any circumstances but for these few days a Washington bureaucrat’s delight.

Roger Chaffee’s attendance was especially significant and attested to the astronauts’ growing awareness of the importance of these trips as well as to Roger’s personal interest. Usually astronauts who would soon receive flight assignments could not take time off to attend to business other than that directly related to their flights, and there definitely was no geology to be done on Apollo 1. Roger enjoyed the training and was becoming an able field geolo­gist. I’m sure he hoped word of his new skills would get back to Deke Slayton and Al Shepard and put him in line for future Moon missions.

I told him I intended to submit my application for the next scientist – astronaut selection and hoped I would soon join him in the astronaut corps. Neither Roger’s flight nor my selection came to pass; less than two years later Roger died tragically in the Apollo 1 fire along with his two crewmates Virgil ‘‘Gus’’ Grissom and Edward White. Their deaths led directly to a major re­evaluation of how NASA was preparing for the Apollo missions, however, and the changes in the way NASA would do business ultimately ensured the pro­gram’s success.

Here is as good a place as any to relate my own experience in attempting to become an astronaut and give some idea of how scientist-astronauts were se­lected. Although I had been a military pilot, as were almost all the astronauts, I didn’t have a lot of jet hours; most of my flight time had been logged on propeller aircraft many years earlier. After working with the astronauts for a year and knowing their flight backgrounds, I could see that it would be virtually impossible for me to qualify in a typical selection process because I lacked current piloting experience. Then I heard that scientist-astronauts might be recruited. In April 1964 NASA asked the National Academy of Sciences to develop procedures for selecting them. Gene Shoemaker had lobbied for such a selection, and before he was diagnosed with Addison’s disease he had been considered a probable top choice when NASA finally got around to agreeing it needed such positions. Even after knowing he would not be selected, Gene continued to lobby, and his efforts, along with those of others in the science community, eventually paid off. I bided my time feeling that my best chance to qualify for the astronaut corps would be through the scientist-astronaut program.

When the call for applications was finally announced in October 1964, I quickly obtained the packet with the paperwork to be completed. It listed standards for such qualifications as age, height, and educational background.

Height! Maximum allowed height was six feet. I was six feet one. The age limit excluded anyone born before August 1, 1930. I was nine months overage. I made a few calls to see if these requirements were inflexible and found that they were. The height restriction was based on the dimensions of the Gemini cap­sules and the Apollo equipment then under design, which would not comfort­ably accommodate anyone over six feet. Greatly disappointed, I wrote to the National Academy of Sciences, the initial screening hurdle, to tell them I was interested but was disqualified because of my age and height, and that I hoped these restrictions might one day be changed so that I and others in my predica­ment could apply.

The good news about this first scientist-astronaut selection was that Jack Schmitt, then working on projects we were sponsoring at Flagstaff, made it all the way through, and he and five others became the first of this special group. Suddenly we were to have a strong advocate in Houston, someone who saw eye to eye with our concerns; but we would have to wait a year for his help while he trained to be a pilot.

I had written to the Academy with deliberate forethought. I felt sure there would be other scientist-astronaut selections. Our post-Apollo planning at that time called for extensive scientific experiments on the lunar surface, and quali­fied scientists would have to perform them to satisfy the scientific community. George Mueller had testified before Congress on these plans, and I knew he supported the need for additional scientist-astronauts. My letter, I hoped, would be retrieved at the next selection, showing my long-term interest in the program and perhaps influencing the selection criteria.

To give myself a better chance in the next selection, whenever it might be, I decided to apply for a pilot slot in one of the Navy Ready Reserve squadrons at nearby Andrews Air Force Base. My last flying experience had been with a navy reserve squadron in Denver while attending graduate school. No pilot openings were available at Andrews in 1964, so I joined an intelligence unit drilling once a month to get back in the Ready Reserve flow and learn through the grapevine where pilot assignments might be found.

This contact soon turned up a vacancy at Lakehurst Naval Air Station, and I quickly transferred to VS-751, an antisubmarine squadron, to resume flying after a seven-year layoff. A year and a half later, with new flying time under my belt, I persuaded a fighter squadron commander at Andrews who needed pilots to have me transferred, and I began the transition to the F8U Crusader. But the navy got wind of this behind the scenes activity; needing antisubmarine – qualified pilots, it rescinded my transfer and assigned me to VS-661 at Andrews. Although I was disappointed (I was looking forward to flying the Crusader, one of the navy’s best-ever fighters), the transfer had one redeeming factor. I would now fly out of Andrews and save the long monthly commute to Lakehurst. And at least I was flying and could hope that this would be a plus in the next selection.

In September 1966 the National Academy of Sciences announced the second scientist-astronaut selection. Accompanying the press release was a short state­ment by Gene Shoemaker, who would be chairman of the Academy’s selection panel: “Scientific investigations from manned space platforms and direct obser­vations on the Moon will initiate a new phase in man’s quest for knowledge. While such missions call for daring and courage of a rare kind, for the scientist they will also represent a unique adventure of the mind, requiring maturity and judgment of a high order.’’ Who could resist such a challenge? I thought that, with Gene as chairman and knowing several other members of his panel, I would have a real chance. It was rumored that this would be a larger class than the previous group of six, thus improving my odds. The Academy had been somewhat disappointed by the number of applications received for the first selection, although the six chosen had excellent qualifications, and thus the selection criteria were a little more relaxed the second time. The age and height limitations had not been changed, but this time the press release stated that “exceptions to any of the. . . requirements will be allowed in outstanding cases.’’ Perhaps now I had a chance. Could I qualify as an “outstanding case’’?

My application must have been one of the first received. As I remember, almost five thousand applications were screened for this second selection. Evi­dently there had been enough good publicity about the Apollo program in the interim to encourage many young scientists to want to be a part of it. About two hundred were selected for the next phase of physical and psychological exam­inations; I made the cut. We were divided into small groups and sent to the Air Force School of Aerospace Medicine at Brooks Air Force Base in San Antonio, where all astronaut candidates were screened.

We endured a week of prodding, blood work, and spinning, IQ, and many other tests, some of which were vividly shown in the movie The Right Stuff, though not with the same comic detail. (For a more complete account of what we experienced, read Mike Collins’s book Carrying The Fire.) While I was tilted upside down with my stomach filled with a barium solution, they discovered that I had a slight hiatal hernia; the muscles in my esophagus couldn’t hold all of the solution in my stomach. Because it was apparently a minor ailment and because, I assume, the other test results were good, I was sent to the Walter Reed Medical Center in Washington, D. C., for a second opinion. The examination at Walter Reed went well, and the examining doctor wrote a letter to NASA saying he did not consider the diagnosis disqualifying—that at the worst I might have to take an antacid to relieve any discomfort I might feel in zero gravity.

Where did this leave me? I couldn’t be sure, but I did have enough experi­ence to know that astronaut selections were secretive. I knew Deke Slayton and Al Shepard were involved, but I didn’t know who else. By this time I was acquainted with all the astronauts, including Al and Deke, but I wasn’t sure whether this was good or bad. I had been on field trips with them, from time to time I was invited to brief the astronauts on the plans for post-Apollo missions, and I was often in the astronaut office building to visit Jack Schmitt and other astronauts as well as the Crew Systems staff. I felt I had a good relationship with them, but perhaps my differences with some MSC managers might hamper my selection. In June I received the call I had been hoping for. I had made the final cut and was invited to Houston for the last interviews before a selection was made.

In June 1967 twenty-one candidates made this final visit. A few of them I knew from my week in San Antonio. Their backgrounds included almost all scientific disciplines, but as I read the list I saw I was the lone geologist, along with one geophysicist. Only two earth scientists! Most of the post-Apollo sci­ence activities we were planning had some earth science connection; I thought my selection was in the bag. The first scheduled activity after checking in was a ride in a T-38, the astronauts’ aircraft of choice, based at Ellington Air Force Base. This was a piece of cake. I flew the plane from the front seat with a NASA pilot (perhaps evaluator?) in the back seat. I did some simple maneuvers and a few snap rolls and generally showed off my flying skills. From what I read in the brief bios of the other candidates, I believed I was the only one with experience as a jet pilot. If this was a test, I must have passed. Next we took a ride in the MSC centrifuge; as I remember, they spun us up to about six gs while we performed a few simple exercises of hitting some light switches. Not a problem, and I suspect some of our future bosses were looking on through closed-circuit television to see how we did on the nearest thing to a stressful test.

After a few other briefings came the interview. I recall only four people in the room: Al, Deke, Bill Hess, and Charles Berry, who was head of the medical sciences office—‘‘the astronauts’ doctor.’’ All the questions were rather innocu­ous. Berry asked about the hiatal hernia, and since I had seen the Walter Reed report I told him that I hadn’t even known I had it until the test and that I didn’t think it would cause any trouble. The only question that stands out in my mind was the one Deke asked: ‘‘Don’t you think you’re too old to be an astronaut?’’ I was thirty-seven at the time and not the oldest of the final twenty-one candi­dates, but I knew I was over the advertised age allowance, so I had done a little homework. I answered, ‘‘I don’t think so; after all, I’m younger than Wally Schirra, and he’s still flying.’’ This brought a big laugh from all four inquisitors. Considering that Walter Schirra, then forty-three, was the only astronaut from the original seven to fly in all three programs—Mercury, Gemini, and Apollo— my answer was evidently on the mark. That ended the interview, and Al said he would give me a call. I thought my selection was now only a formality. That afternoon I did some preliminary house hunting in the neighborhoods around NASA.

In August Al called. ‘‘Don,’’ he said, ‘‘I’m sorry to tell you you weren’t selected.’’ We talked for a few more minutes, and I’m sure he realized my disappointment. They had chosen eleven for the scientist-astronaut class of 1967, including the geophysicist Anthony England, the only other earth scien­tist. I didn’t ask why I wasn’t selected; I was sure he wouldn’t give me any specifics. I rationalized that it was a combination of things. My hiatal hernia (they didn’t have to take any chances on its causing a problem); my seniority (from a government classification standpoint I would have been senior to most of the astronauts selected earlier); my pilot background, which may have been seen as a negative (I would have been the only one they didn’t have to send to pilot training, and that might have made me an apple among all the oranges. What would they do with me during the year the others were in training?) Finally, they might have received some negative comments from MSC man­agers I had disagreed with in years past.

Alan Shepard died recently, so I won’t get a chance to ask him why I wasn’t chosen. Perhaps he would have told me, perhaps not; most probably, after so many years he wouldn’t even have remembered. In any case, the rejection probably did those of us not selected a favor from a career standpoint. Within three years the post-Apollo missions, the prime reason for the selection, were canceled, and none of the class of 1967 flew on a space mission for fifteen years; Joseph Allen was the first from this class to fly as a mission specialist, on shuttle flight STS-4. A few retired or left NASA before taking part in any NASA mis­sions, and several, like Joe and Story Musgrave, made major contributions to NASA programs.

Returning to training and simulations, geological field training for the astro­nauts became more and more realistic and intensive as the date for the first landing came closer. By 1966 all the astronauts had had some level of both classroom and field training. Those in the first three groups selected had the most extensive geological training. Since no one knew who would ultimately be selected for the landing missions, we tried to have them all at as high a level of competence as possible within the time available. Many noted geologists volun­teered to assist in the training; some stayed on to become members of the Apollo Field Geology Team and worked with the astronauts until the last mis­sion, Apollo 17, was safely home. Lee Silver, Richard Jahns, Aaron Waters, Dallas Peck, and William Muehlberger come immediately to mind as volunteers who devoted a significant part of their professional careers to these efforts. Many others made important contributions to astronaut training, including many geologists on the staff at MSC.

I was able to take part in several field geology training trips, and those I attended were all memorable. A specially arranged visit to the Pinacate volcanic fields in Sonora, Mexico, just over the border from Arizona, had a somewhat different purpose. This trip took place in late summer 1966. The Pinacate area includes an interesting set of volcanic craters formed by the explosive release of superheated underground water; craters of this type have their own geologic name—maars. From the air they have an uncanny resemblance to some lunar craters: their rims are only slightly raised, the craters themselves are symmetri­cal, and many are relatively shallow. Some of those at the Pinacate are quite small, a few hundred feet across, and two are very large, the largest being over one mile in diameter. The area where they occur is desolate and isolated, a perfect place to take a high profile group like astronauts, where no one would disturb their training. (It was definitely a place where reporters would not go, for there were no amenities of any kind.)

The Pinacate became one of the favorite training sites, and most of the astronauts made a visit at one time or another. This visit was without astro­nauts; its purpose was to educate my bosses, Phil Culbertson, who had replaced Tom Evans in August 1965, and his boss, E. Z. Gray. Since we were still looking for new training sites for the post-Apollo missions, I thought it was important to show E. Z. and Phil how we would use such sites and what benefits could derive from good terrestrial analogues like the Pinacate. I had arranged with Gene Shoemaker and Al Chidester to conduct the trip as if it were an astronaut training trip, with Phil and E. Z. being treated, in a manner of speaking, as the training subjects.

For both of them it would be a real eye opener; we would camp out in tents for two days in the middle of nowhere, something they had seldom experi­enced. We all flew in to Phoenix and were met by the USGS staffers who would support the trip. Then in a caravan of four or five trucks we turned south on Route 85 with a first stop at Ajo. At that time Ajo was a copper company town with a company store that sold provisions at a discount; the USGS guys always knew how to save a buck. Among other food, we bought frozen T-bone steaks to grill over an open fire the first night; with no refrigeration, we had to cook them that day, and by the time we made camp we expected they would be thawed. From Ajo south, Route 85 takes you through Organ Pipe Cactus National Monument, a unique desert habitat with numerous large saguaro cacti standing like statues along the highway and stretching off into the distance in all direc­tions. This was the ‘‘green desert,’’ with all kinds of unusual plant life including mesquite, palo verde, cholla, and other thorny stands of wicked-looking cactus that I had first seen when we visited Fort Huachuca.

We crossed the border at Lukeville and turned west on Mexico Highway 2. Almost immediately the landscape changed dramatically, becoming much more barren and arid with only a few scattered houses along the road out of Sonoita, the small Mexican town opposite Lukeville. After a few miles we turned off on a dirt road and continued south; the dirt road turned into two tire tracks, and finally we drove in and out of the dry arroyos, gaining a little elevation, and arrived at the volcanic fields about three in the afternoon. While the USGS support team set up camp, we walked over to the rim of Elegante Crater for our first look at the next day’s simulated training site. Elegante Crater is impressive. Over five thousand feet in diameter and eight hundred feet deep, it was not unlike Meteor Crater in many respects, except there were no large blocks of ejecta around the rim and few blocks or large boulders in the interior. The crater looked as though it had been scooped out of the desert by a large spoon, and whatever had been in the center had disappeared. These craters normally constituted a difficult test for the astronauts to interpret and describe so that the accompanying geology staff, acting out the role of a support team back on Earth, could develop a reasonable geologic map based on the astro­nauts’ descriptions.

By the time we returned to camp the tents were all set up and a campfire was lit. Gordon Swann and I went back to the pickup for the frozen steaks and lifted the cardboard carton to carry them over to the cook. They had thawed, the carton had turned to mush, and the thirty or so steaks fell through the bottom onto the sandy soil. What a mess. With a carefully rationed supply of drinking water to last the two days, we could spare only a little to wash off the steaks, so they were still crusted with sand when they finally hit the grill. E. Z. and Phil, along with the rest of the crew, were treated to a new dinner sensation: steak that wore your teeth down if you bothered to chew. I could tell E. Z. wasn’t enjoying his outing—not the best way to impress the bosses with how well organized we were on astronaut training trips. Around the campfire that night the veterans of this type of trip told tales of previous visits to the Pinacate and described some of the exploits they had been party to. Some of the astronauts were enthusiastic card players, and apparently a few exciting card games on past visits had gone on into the wee hours, affecting their next day’s concentration and ability to absorb some rather detailed geological lectures. As we knew, not all the astro­nauts took the field training seriously.

We were a much more sedate group than some in the past, except that a couple of USGS staffers had brought the makings for powerful after-dinner drinks. By the time the storytelling was in full swing, several in the cast were oblivious to the heat and sand. Those of us who were not imbibing heavily decided to call it a day, and along with Phil and E. Z. we crawled into our tents. With fewer seniors around the campfire to dampen the storytelling, the talk grew louder and louder, punctuated from time to time by the equivalent of an Arizona rebel yell. Finally E. Z. couldn’t take the noise any longer. He jumped out of his tent and threatened to cut off all USGS support if they didn’t imme­diately shut up and go to bed. This got their attention; the noise decreased to a low rumble and then silence. When we finally fell asleep, all we could hear was the buzzing of the night insects.

The next morning, up with the sun, we were gathered around the fire awaiting breakfast and the first geology lecture when we noticed that two staff members were missing. We searched around the campsite and couldn’t find the midnight revelers. We were getting worried; rattlesnakes, scorpions, and gray wolves inhabited this area, and there was even an occasional panther. Finally we found one of them asleep in a truck cab, and the other turned up several hundred feet from the camp, lying near a clump of cholla, slightly the worse for wear with his shirt torn and a little bloody. Thus was added another chapter of tall tales for future astronaut training trips. But for E. Z. Gray it was the last straw; he cut his visit short and was taken back to Phoenix that afternoon. By the time I got back to Washington he had calmed down, and we continued to support USGS in all its work. Training trips to the Pinacate were considered highly successful, and on missions to the Moon some of the astronauts would comment on how much the Moon’s surface looked like their memory of the Pinacate.

Mission simulations for crews assigned to specific Apollo lunar landing flights had a somewhat different aspect. For these exercises the two astronauts assigned to the lunar module would be involved, often with their backup crew and sometimes with the command and service module crew member, depend­ing on the objective of the simulation. This meant a support crew of dozens. In addition to the astronauts, lecturers, and technicians, the ever present MSC photographers would be milling around snapping pictures from all angles. Walt Cunningham’s simulation at Bend, Oregon, was an intimate gathering (with the exception of the press that was present) compared with these later simulations. As we approached the flight date, simulations would progress from casual dress at analogue field sites to full suited simulations at MSC or KSC, with some of the latter attempting to follow projected lunar timelines as closely as possible.

As principal investigators were identified for each of the science experi­ments, they would also attend from time to time, along with the contractors building the equipment, so they could observe how the astronauts deployed or operated their instruments. At times the simulations would result in changes to accommodate the astronauts’ ideas on how to improve their interaction with the particular experiment; but whenever possible the astronauts attempted to adjust to the idiosyncrasies of the experiment and achieve the best results for the PIs.

By this point in the training (crews being selected for specific missions), the simulation sites included an MSC high-bay building, the ‘‘back lot’’ at MSC, a small outdoor site at KSC, and a few special analogue sites scattered around the country, chosen to be most like what the astronauts would find on the Moon. The MSC ‘‘back lot’’ or ‘‘rock pile’’ was a few acres of simulated lunar terrain with an LM mockup in the center. The surface was covered with gravel and sand and salted with various types of rocks. A smaller simulated outdoor lunar surface was built at KSC, primarily as a convenience for the astronauts, who spent more and more time there as their launch date approached. The KSC site was often unusable because the ‘‘craters’’ would fill with water at high tide (very unmoonlike), but this site permitted last-minute reviews of specific tasks that may have been added or modified since the previous simulations at MSC. The KSC outdoor site did not include an LM mock-up, so it could support only limited types of simulations. However, there was an indoor site that did include an LM simulator. The KSC simulations were usually conducted in pressure suits to be as authentic as possible. Equipment provided was spare flight article hardware or the closest copy we could obtain.

One of the special analogue sites was near Sunset Crater, a few miles north­east of Flagstaff. Calling it an analogue is a bit of a misnomer, because it was in fact the closest copy of a moonscape that existed anywhere on Earth. Some of the staff at Flagstaff hit on the idea of duplicating the lunar surface as seen in one of Lunar Orbiter’s pictures. They carefully analyzed the selected frame, measuring the diameter and depth of all the small craters and interpreting the history of this small piece of the lunar surface by determining the relative age of each crater based on how the ejecta layers overlay each other. After these cal­culations were made, Norman ‘‘Red’’ Bailey and Hans Ackerman, two Astro – geology staffers, laid out a grid of fertilizer bags on a ten-acre volcanic ash fall south of Sunset Crater. When the fertilizer and fuel oil explosive was detonated, the Orbiter photo was recreated. Not only were the bags arranged according to the explosive force they would generate to create the proper size craters in the correct locations, but they also were timed to go off in the sequence that would provide the correct ejecta layers observed on the real lunar surface. It was a roaring success in all respects, and the creation day was delayed until I was able to witness it on one of my frequent trips to Flagstaff. A movie was made of the explosions, and it was great fun to replay it for visitors who came to watch the astronauts training at the site; each new crater erupted in sequence, in slow motion, and the fine ash flew skyward in great dark jets.

This site, and two additional sites formed in the same manner, became the last tests for the astronauts, requiring them to use all the observational skills they had gained. As they walked or drove around on the closest thing to the Moon they would see until they actually landed there, they described it to the backroom crew so that a geologic map could be made. After completing the exercise, they would review their observations with their instructors to correct any misinterpretations they might have made. All the astronauts from Apollo 13 onward trained at these sites, and I always thought it was one of the best simulations they were involved in, since it was the most complete test of their skills at observation and description.

A drawback with all the pressure suit simulations was that we could not replicate the one-sixth gravity field they would experience on the Moon. In some sessions we tried to simulate the low lunar gravity by using two types of simulators and specially rigged harnesses that partially suspended the test sub­ject and reduced his weight to one-sixth of his Earth weight. These simulations were usually not very satisfactory because the complicated harness setup would reduce only the astronaut’s apparent weight, not the weight of the equipment he was working with. But some of these tests provided important insights, since the mass of the equipment was accurate and the astronauts got a feel for this unique combination of forces. The NASA airplane, normally used to simulate low or zero gravity, also was a poor substitute because of the short duration of each flight parabola. Neutral buoyancy simulations (held in a tank the size of a swimming pool)—a much better way to simulate low gravity environments and the standard way to train for today’s shuttle missions—were in their infancy. They were used for simulating the zero gravity parts of the missions, but not for lunar surface tasks.

In addition to simulating the geologic tasks they would carry out, the astro­nauts simulated the deployment of the Apollo Lunar Surface Experiments Package and the use of all the other equipment and experiments they would carry on the mission. For the final three missions the important equipment additions were the lunar roving vehicle and the lunar drill. The LRV’s deploy­ment from its stowed position on the LM landing stage became a critical part of the timeline. To accomplish all the tasks planned for the extended-staytime missions, the astronauts had to get the LRV functioning as quickly as possible. This meant removing it from the LM stowage bay and setting it on the surface while simultaneously unfolding the wheels tucked beneath the frame, erecting the TV and communication antennas, and finally checking the drive system to be sure it had survived the long journey. A clever but complicated system of cables, springs, and hinges was designed for the LM and LRV.

Once they were sure the LRV was operating correctly, they would load it with the other equipment and experiments that depended on the LRV for their operation. LRV deployment was rehearsed over and over again to reduce the time it took and try to ensure success. During the training sessions the MSC and KSC staffs would introduce hang-ups in the deployment of the LRV and other equipment to see if the astronauts could overcome such adversity. They soon became adept at doing this and foreseeing problems.

Another important task to simulate was getting the loaded lunar sample return containers back into the LM from the lunar surface. This maneuver tested the ingenuity of the MSC engineers because the astronauts could not carry the bulky containers up the LM ladder. They devised a pulley system. One astronaut would kneel in the LM hatch while the other stayed on the surface to hitch the containers to the pulley cables and slowly pull them up to the waiting astronaut. Although it was a relatively straightforward solution, the cable sys­tem tangled easily, so it took many hours of practice to rig the pulleys and coordinate the two astronauts’ actions. Lending urgency to these ‘‘rock box’’ simulations was the knowledge that of all their tasks this was the most impor – tant—the harvest of Moon rocks and soil. If for some reason the sample con­tainers were left behind, the mission would be deemed a failure. This would be especially true for the final missions, which would include samples from loca­tions far from the lunar equator and precious cores collected from below the lunar surface by the lunar drill.

By mid-1967, detailed training and simulation schedules were set up for each of the lunar landing missions.10 Starting forty-four weeks before their scheduled launch date, the astronauts would follow a tight schedule designed to cover all aspects of the missions. Almost 2,200 hours of training and briefings were crammed into their workdays at both MSC and KSC. Some required the presence of all three astronauts, others called for the CSM pilot alone, or just the two Moon-landing astronauts. This constituted a scheduled fifty-hour workweek for each of the three astronauts and the backup crew, with untold extra hours of unscheduled time. They underwent a minimum of 5 hours a week of physical training, 6 hours a week of flying time, 5 hours a week of Apollo flight plan reviews, and 25 hours of flight-suit fit checks, 196 hours of spacecraft tests, 20 hours reviewing stowage procedures for both the CSM and the LM, 40 hours of planetarium exercises to ensure that the crew could use celestial navigation to update their programed navigation system in case of several possible failures, 10 hours of egress training to cover water recovery from the CSM after splashdown, 269 hours of briefings and simulations for science operations, and many other types of training. The 269 hours of science training was one of the largest time allocations, and it was jealously guarded by those of us involved in providing the science payloads, since the other side of the NASA house—the engineers, flight controllers, and other critical partici­pants in launch preparations—would try to preempt some of this time for their own use. But in spite of this constant demand for more astronaut time to attend to nonscience matters, Deke and Al stuck to the schedule, and we were seldom shortchanged. After being named commander for Apollo 14, and while involved firsthand in the training for his mission, Al became a strong supporter for the science team’s training requirements for the final three missions.

When the contract was signed to build the LRV for the last three missions, Rutledge ‘‘Putty’’ Mills, our vehicle guru at Flagstaff, was charged with building a training vehicle that would approximate the LRV configuration so that we could continue to do mission planning and simulations at Flagstaff. (The flight version of the LRV could not be used in terrestrial simulations because it was designed to operate in lunar gravity. It would have collapsed under the astro­nauts’ Earth weight.) An LRV simulator that could be used in Earth’s gravity was not due from the contractor for some months, and we wanted to get an early start on our simulations. Putty did his usual innovative job of construct­ing a vehicle from odds and ends and his fertile imagination. We named it ‘‘Grover the Rover,’’ for one-g rover, and it was ready for testing by the end of June 1970, just six months after Boeing was given the final LRV specifications. At the end of August we conducted a full-scale test, with astronauts participat­ing as well as others. Astronauts in attendance were John Young, Charles Duke, Tony England, Gerald ‘‘Jerry’’ Carr, William Pogue, and Fred Haise.

The test was scheduled to be conducted at the Cinder Lake Crater Field Number 1, but most of the driving over the next four days took place at a vacant lot near the USGS building in Flagstaff. The astronauts present operated the Grover, as did engineers from MSC, MSFC, and NASA headquarters. Putty had built the Grover to run on electric motors like the real LRV, and he had three battery packs available to recharge so we could have more or less continuous operation. At full throttle the Grover could make seven miles an hour carrying two passengers, similar to what we could expect of the LRV on the lunar surface. Mock-ups of some of the tools were stowed on a pallet on the vehicle, the way we anticipated they would be carried on the Moon, although a final stowage configuration for the LRV had not been decided. At the end of the test, all agreed that the Grover would be a valuable addition to future mission simulations, especially when Putty had a chance to add refinements such as a navigation system and additional mock-ups for the lunar communications relay unit, TV, and other equipment the LRV was scheduled to carry.11 Even­tually we obtained a fully functional spare LCRU for our simulations.

A site selected for the simulations conducted toward the end of crew training for the final missions was on the island of Hawaii. Despite the prevailing view that most lunar features were the result of impact processes, all the astronauts had visited Hawaii early in their geologic training to study the wealth of lunar – like features created by the many active or semiactive volcanoes. Simulations for specific missions were a different matter, more like a final exam. We chose several locations on the island to represent geological situations similar to those the crew might encounter on the Moon. Typifying the Hawaiian simulations, the Apollo 17 crew spent the first four days visiting these sites, then had a day of rest. Dallas Peck, a noted volcanologist who had spent a number of years in Hawaii studying the island’s geology, acted as coordinator and principal lec­turer. The final three days were spent at Kahuku, Hualalai, and the volcanic ash wastelands at the crest of Mauna Kea (elevation 13,796 feet), chosen to repre­sent what astronauts Gene Cernan and Jack Schmitt might find at their desig­nated lunar landing site, the Taurus-Littrow Valley.

At Mauna Kea the staff had prepared a series of traverses around the vol­cano’s summit that would approximate those the crew would follow on the lunar surface. Sampling and description stations had been designated at inter­vals replicating as closely as possible the Taurus-Littrow timeline that had al­ready been carefully plotted by the Field Geology Team for the actual mission. All the surface equipment the crew would deploy or operate, except for ALSEP, was transported to the top of the crater, including a simulated version of the LRV. Putty Mills had modified a local jeep to use as a simulated LRV, a cheaper and less sophisticated version of the Grover and other LRV training vehicles. It also avoided the expense of transporting one of these trainers from the main­land to Hawaii. He had removed most of the jeep’s body and engine so that the astronauts were sitting on open seats on the frame and could climb on and off easily. He had also added racks for their tools and sample bags and a mount for their communication antenna, similar to the stowage on the real LRV.

During this training exercise most of us lived in motels on the coast, either in Hilo or in Kailua-Kona, commuting the thirty to forty-five miles a day to the training sites. Some of the USGS staffers lived closer in an army base and kept most of the equipment we would use each day there. Cernan and Schmitt wore street clothes for these simulations; it would have been too costly and time consuming to try to conduct them in pressure suits this far from Houston. To add some mission reality they wore backpacks similar to the portable life – support system, but with battery power only for voice communication back to our simulated Science Support Room out of sight of the traverses.

Bill Muehlberger, the Field Geology Team PI appointed for Apollo 16 and Apollo 17, was in charge of this trip. He brought several members of his team including George Ulrich, Gerry Schaber, and Dale Jackson. Scientist-astronaut Robert Parker was also on hand, since he had been designated mission scientist and the prime capsule communicator during the periods of extravehicular activity. Muehlberger and his team would man the rudimentary SSR, connected to the astronauts only by radio, plotting their progress as they drove around the summit and communicating through Parker as they would during the actual mission. The Field Geology Team, through trial and error on earlier missions, had devised procedures to assist the astronauts if something unexpected hap­pened or to respond to any questions they might have, and these procedures were also practiced.

Those of us not directly involved in the backroom simulation would follow Cernan and Schmitt from a distance as they drove from station to station, making note of how everything fit together—or didn’t, as the case might be. At the end of the exercise, Muehlberger and his team retraced the traverses with Cernan and Schmitt, reviewing how they interpreted their voice reports, cor­recting their map, and then suggesting ways to improve the crew’s descriptions to produce a better interpretation of what they actually saw.

With the first scientist-astronaut geologist in the crew and a highly moti­vated and well-trained commander, we didn’t expect there would be much need for this type of support, but as with all things NASA, we were going to be prepared. All in all, this Hawaii simulation was about as good as we could get in obtaining a high fidelity rehearsal before the real mission was under way.

We conducted one week of intensive, almost uninterrupted training for both

the crew and the Field Geology Team. Apollo 17 would be the last mission, and Muehlberger was determined that it would be the best if he had anything to do with the training and simulations. In just five months it would be the real thing. A final reward for our efforts had become a tradition. On the last night of these trips, a dinner was held at Teshima’s, a lovely Japanese restaurant high on a hill overlooking the ocean, with Mrs. Teshima providing a royal welcome and a special menu. It was a night of storytelling, practical jokes, and reminiscing, a dinner that all who attended will long remember.

Studying the Moon from Orbit

Although the Ranger and Surveyor missions had sent back many close-up views of the lunar surface, they were never intended to provide all the photographs we would need to select the Apollo landing sites. That was to be the job of Lunar Orbiter. Conceived in 1963, its objective was to obtain detailed photographs of the whole Apollo landing zone. We needed high resolution in order to pick areas free of large boulders or small craters that would be a hazard to the astronauts guiding the lunar module to a safe landing. Obstructions of this size could not be seen on photographs taken from Earth, even by the largest tele­scopes. The Lunar Orbiter program was managed by the Office of Space Science (later the Office of Space Science and Applications), but the photographic design requirements were dictated by the Office of Manned Space Flight and in particular the engineers at the Manned Spacecraft Center. Langley Research Center (LaRC) was selected to be the day-to-day manager, and the request for proposal was released by LaRC. The RFP called for building six to eight orbit – ers; it was possible that the final ones in the series would include other experi­ments in addition to cameras. OSSA released an announcement of flight oppor­tunities to solicit experiments for these last missions and received over one hundred proposals or inquiries.

The competition to build the spacecraft and cameras was won by the Boeing Company as the prime contractor, supported by two major subcontractors, RCA and Eastman Kodak. Langley’s program manager, Clifford Nelson, put together a superb team to oversee the program; many years later, when NASA management called for a review of lessons learned from all the completed programs, Lunar Orbiter was judged the best managed. If for some reason it had not been successful, the entire Apollo project would have been in jeopardy or, at the very least, delayed beyond the date President Kennedy had called for. Lunar Orbiter was successful far beyond our hopes based on our experience with Ranger and Surveyor. Lunar Orbiter 1, which flew in August 1966, did not perform completely to specifications, but it returned a total of 422 medium and high resolution photographs of potential lunar equatorial landing sites as well as some photographs of the Moon’s farside. After correction of the problem that degraded some of the first mission’s photographs, Orbiter 2 and Orbiter 3 were so effective that all the Apollo landing site photographic requirements were completed; the engineers and mission planners had enough photographs in hand to permit detailed landing site analysis, and they released the final two spacecraft for science and site selection for potential post-Apollo missions. (The last three Lunar Orbiters were eventually canceled, and the experiments solicited for those missions were put on the shelf to be resurrected later.)

The first three spacecraft had concentrated primarily on photographing the nearside equatorial zone, where the upcoming Apollo landing sites would be. Lunar Orbiter 4 expanded the coverage on the nearside, including many of our high priority post-Apollo exploration sites. The final mission, Lunar Orbiter 5, completed the coverage of the poorly known farside. By the time Lunar Orbiter 5 snapped its last picture, the five Lunar Orbiters had sent back 1,950 pictures of the Moon covering most of the lunar surface, nearside and farside. The resolution of these photographs ranged from approximately sixty-five meters to five hundred meters, although much higher resolution photographs of the potential Apollo landing sites were taken on the first three missions. To obtain this higher resolution (two meters), the first three missions took their photo­graphs at lower orbital altitudes than the final two.

Thus Lunar Orbiter equaled the best Earth-based photographs, and it bet­tered many of them by a factor of 250. Only a small area of the Moon was covered by the high resolution photographs, but the coverage had been judi­ciously distributed by the planning teams. An added benefit was that by closely tracking the spacecraft’s orbits, we were able to map the Moon’s gravity field at a resolution not achievable from Earth.

Both the Falmouth and Santa Cruz summer conferences devoted consider­able thought to recommending experiments that could be done in lunar orbit to complement the study of the Moon from the lunar surface as part of the comprehensive, post-Apollo exploration program. In 1964 and 1965 Peter Badgley had attempted to interest NASA management in a remote sensing program to be conducted in Earth and lunar orbit, and eventually a program titled Lunar Mapping and Survey System was initiated.1 This program, designed to use Apollo hardware, was canceled in early 1968 in a cost-cutting move.

But the recommendations from the summer conferences did not die. In March 1968, ignoring the just announced program termination, Sam Phillips sent a memo to Bob Gilruth requesting that MSC look into providing scientific and operational photography during the landing missions.2 With planning proceeding for the final missions, and following up on the Phillips’s request, Lee Scherer sent Bill Hess a memo in early May 1968 asking that MSC begin to study how to integrate experiments into the command and service module to take advantage of the longer staytime in lunar orbit. Hess agreed, prompting our office to write a memo for Phillips’s signature asking MSC to expand the study he had requested in March to identify other orbital experiments that would take advantage of the ‘‘overall CSM science potentialities.’’3 This memo resulted in MSC’s adding $100,000 to its Martin Marietta Apollo Applications Program integration contract and marked the beginning of a program to de­velop a suite of sensors that would be flown in the CSM.

While this analysis was under way, OSSA dusted off the experiments that had been submitted earlier for Lunar Orbiter and began to assemble the ra­tionale for including different suites of cameras and sensors that could fit into the CSM. George Esenwein, who had been the headquarters project officer for the Apollo command and service module mechanical systems, transferred to our office at this time and was put in charge of the orbital science and pho­tographic team. Floyd Roberson was named program scientist, and David Win­terhalter was program engineer. Noel Hinners, at Bellcomm, assigned several members of his staff to work with this team, notably Farouk El Baz and Jim Head, both of whom had played prominent roles in analyzing Lunar Orbiter photographs and recommending targets for photography on Orbiter 4 and Orbiter 5.

As an extension of these studies, Esenwein’s team, working with MSC, deter­mined that it would be possible to include in a service module (SM) bay a small subsatellite that could be left in lunar orbit, and an announcement of flight opportunities was released soliciting experiments that could utilize the sub­satellite. In April 1969 OSSA and its advisory panels reevaluated the Lunar Orbiter proposals, and the new proposals to place experiments on the sub­satellite, and selected a final suite of experiments.4 In June OMSF directed MSC to proceed with the modifications of the CSM and to procure the experiments. Eventually the science payload carried in the command and service module, including cameras, experiments, and the subsatellite, totaled almost 1,200 pounds. Most of the experiments were housed in one quadrant of the service module in what was named the scientific instrument module (SIM), and a few were carried in the command module (CM).

For the experiments that did not send their data back by telemetry but recorded them on film or in some other form, the film and data would have to be retrieved by the CM pilot during extravehicular activity. After much debate concerning the safety of the CM pilot during the retrieval operations, it was finally agreed to schedule this EVA after leaving lunar orbit, when the astro­nauts were safely on their way back to Earth. Imagine floating outside your spacecraft somewhere between the Moon and Earth attached by an umbilical cable and a slender wire! The three CM pilots who carried out this risky maneuver would all comment on the strange sensation of seeing the Earth from so far away while floating in space.

Starting with the flight of Apollo 8 at Christmas 1968, the astronauts began making their contributions to studying the Moon from lunar orbit. Armed with the ever present hand-held Hasselblad cameras, the crew of Apollo 8 and all the crews that followed (except Apollo 9, which remained in Earth orbit) took pictures of the Moon from various altitudes above the lunar surface. Many of the photographs taken during the early missions were meant to improve our understanding of future landing sites by augmenting the Lunar Orbiter photo­graphs. Apollo 12, as an example, took 142 multispectral photographs of the designated Apollo 13 landing site, Fra Mauro, and other equatorial sites. These photographs were used to help decipher the geology and to improve the pro­ductivity of the astronauts after they landed by identifying sampling sites that probably had different mineralogical compositions. After Apollo 13’s failure, Fra Mauro became the landing site for Apollo 14, and the information obtained from the multispectral photography helped, in a small way, in planning the Apollo 14 surface traverses.5

Apollo 14 carried out a variety of experiments, including photography, while on the way to the Moon, in lunar orbit, and on the return to Earth. Three types of cameras were used: a 16 mm data acquisition camera, Hasselblads, and the Hycon lunar topographic camera. (The Hycon malfunctioned during the mis­sion, but almost two hundred usable photographs were recovered.) These ex­periments included measurements of gegenschein and heiligenschein (rather arcane observations, the former possibly related to Earth-Moon-Sun libration points6 and the latter related to reflected light, which had potential application for the interpretation of the Moon’s fine-scale surface roughness). An S-band transponder experiment provided new information on the Moon’s nearside gravity field by permitting close tracking of the CSM’s orbits and a bistatic radar investigation that yielded information on the lunar crust.7 The final missions, Apollos 15, 16, and 17, had much more extensive orbital science payloads than any of the previous missions.8

Because I was not closely involved with developing the experiments carried in lunar orbit, I will not further describe them or their principal investiga­tors, but for completeness in covering the scientific results of Apollo, in chap­ter 13 I briefly discuss the scientific information returned from some of the experiments.

On to the Moon:. Science Becomes the Focus

On July 16, 1969, along with a multitude of other sightseers (local Civil Defense officials would later estimate one million), my family and I were on hand to watch the launch of Apollo 11. Our Winnebago camper was parked on the shoulder of U. S. Route 1 about five miles north of Kennedy Space Center and the launch site. We had picked our viewing point the night before, feeling lucky to find a spot so close. It had been a madhouse trying to drive near the Cape; no one seemed to care about following normal rules of the road as cars and campers vied for spots and parked wherever they pleased. Local and state police tried to maintain some order, but it was a hopeless job. In the early morning, as launch time approached, we climbed on the roof of our camper to get an unobstructed view, meanwhile listening on the radio to John ‘‘Jack’’ King, ‘‘the voice of Apollo,’’ count down the final seconds.

Old Glory was flying everywhere, and the crowd was in a party mood. The countdown proceeded smoothly, and at 8:32 a. m. the Saturn rocket lifted off ac­companied by loud cheers and many teary eyes, mine included. Beyond a doubt our hearts went with the crew of Apollo 11. This was the second Saturn У launch I had witnessed, but I still wasn’t prepared for the enormous noise and low – frequency reverberations that reached us, even at this distance, in the minute after the Saturn cleared the launch tower. We watched for several minutes as it disappeared to the east, leaving behind a huge plume of white smoke, then we went inside, finished breakfast, and talked about what we had just seen. My sons, only eight and eleven at the time, still vividly recall the excitement of that morning. I was in a hurry to leave because I was due back in Washington in a few days, but we were forced to wait almost an hour before the traffic jam began to move and we were back on the road. Apollo 11 was on its way to the Moon with the first science payloads that men would place on another body in our solar system. If all went as scheduled, Neil Armstrong and Buzz Aldrin would have the honor of making the first direct, close-up studies of how the Moon’s surface looked and how it felt to walk on the Moon in one-sixth gravity. After the landing and takeoff from the Moon, Mike Collins, the command module pilot, would be waiting in lunar orbit to rendezvous with the lunar module, ready to lower his orbit if the ascent stage did not perform as well as planned.

Four nights after the launch, in anticipation of the landing, the Voice of America (VOA) had assembled a team to report on this once in a lifetime adventure for its worldwide audience. Several NASA colleagues, Merle Waugh, John Hammersmith, William Land, and I, were in the Washington studios as ‘‘color commentators” to back up the VOA reporters led by Rhett Turner, who would be reporting from the Manned Spacecraft Center in Houston. We lis­tened anxiously, just like millions of others around the globe, to the exchange between the capsule communicator (CapCom) Charlie Duke and Armstrong and Aldrin in the Eagle as they went through the final maneuvers to land the LM. The excitement of those last few minutes, heightened by the crew’s diffi­culties in selecting their landing site with alarms ringing in their ears and their fuel supply nearing exhaustion, made Armstrong’s announcement ‘‘Houston, Tranquility Base here, the Eagle has landed,’’ almost anticlimactic. We could hear the cheering in the Mission Control Room through Rhett’s microphone, and we in VOA’s Washington studio were yelling and pounding each other on the back too. Although we had worked for years to help achieve this moment, it seemed incredible that we were successful on the first try.

We were primed to discuss the mission in great detail, but as the night unfolded only a few questions were directed our way, and I was never called on to demonstrate my vast insight into things lunar. VOA wasn’t about to share the limelight on this historic occasion. I did, however, receive a card from some friends in Colombia who said they had heard me on VOA. They told me how proud they were of Apollo 11 ’s success and congratulated me on being part of the program. I wondered if some of my former colleagues remembered their skepticism six years earlier when I decided to leave Mobil and join NASA. I certainly did not regret the decision. Our great hopes to follow Apollo with extensive exploration and lunar bases now seemed remote, but important work still lay ahead to make each succeeding mission more scientifically productive.

As the scheduled launch date for Apollo 11 drew closer, NASA management became more and more cautious and conservative. This was especially evident at MSC, where caution was the trademark, but even at NASA headquarters one could sense growing concern about the many uncertainties and dangers that simulations and planning could not make go away. Mueller’s decision to go to ‘‘all up testing’’ had eliminated several test flights that would have provided additional experience, but it was too late to go back and build confidence any further than where we were in July 1969. The only alternative was to schedule a conservative mission profile leaving as much margin for error as possible.

The Early Apollo Scientific Experiments Package (EASEP) that Armstrong and Aldrin would carry on their flight, described in chapter 7, did not represent a complete Apollo Lunar Surface Experiments Package (ALSEP), since both headquarters and MSC feared that the tasks originally planned would be too demanding. EASEP included a solar-powered seismometer and an additional experiment, the Laser Ranging Retro-Reflector (LRRR). The Swiss-sponsored Solar Wind Composition collector would also be deployed, but its scientific value would be degraded because of the short time it would be exposed to the solar wind. Sample collection and photography were scheduled in connection with the crew’s geological study, but they were also reduced in scope from the original plans.

Before the launch, word of changes had reached Congress, some of whose members were already chafing at the expense of the program. These changes had raised questions about the cost of removing the planned equipment from the Apollo 11 mission. On March 13, 1969, just four months before Apollo 11 ’s scheduled launch, the House Subcommittee on Space Science and Applications held a hearing at which a number of questions were asked about the last-minute science payload changes. Chairman Joseph E. Karth (D-Minn.) asked, ‘‘Can we put in the record why the ALSEP is not flying on the Apollo trip as originally planned?’’1

Our office responded four days later with the following explanation:

The goal of the first Apollo mission to the lunar surface is the successful landing and safe return to Earth of the astronauts. The primary objective of the mission is to prove the Apollo system-launch vehicle, spacecraft, space – suits, men, the tracking network, the operational techniques.

The first landing mission represents a large step from orbital operations.

The descent, landing, extravehicular activity (EVA) and ascent from the lunar surface are new operations in a new environment. Our Gemini EVA experi­ence showed that a methodical increase in task complexity was necessary in order to understand and operate in the zero g space environment. The 1/6 g lunar surface environment will be a new experience. We cannot simulate it completely on Earth. We find, for example, that we simply do not have as much metabolic data as we would like in order to predict with high confi­dence, rates in a 1/6 g environment. Only educated guesses are possible on the difficulties the astronaut will have in maneuvering on the surface or the time it will take him to accomplish assigned tasks.

Until recently, the first mission plan called for two periods on the lunar surface (EVAs). During the second period, the crewmen would deploy the Apollo Lunar Surface Experiments Package (ALSEP). This would take place immediately prior to lunar ascent and rendezvous. Because of biomedical unknowns, we are concerned with the degree to which the second EVA would fatigue the crew and adversely affect their performance during the critical ascent and rendezvous phases of the mission.

After extensive review and evaluation, we reached the decision not to have a second EVA on the first landing mission. The ALSEP will be deferred to the second mission. We will make every effort on the first landing to obtain data leading to a firm assessment of the astronaut’s capabilities and limitations on the lunar surface with a view toward increasing, on subsequent landings, the percentage of EVA time available for scientific investigations. Deployment of the ALSEP on the second mission is planned as a primary objective.

Our answers to other questions raised by the subcommittee included an esti­mate of $5 million to modify the ALSEP seismometer to the EASEP configura­tion. (This number differs from the contract cost of $3.7 million discussed in chapter 7 because it includes other costs associated with the EASEP, such as integration and training, that were not part of the Bendix contract.)

Left out of the response was another concern, the performance of the LM during the first landing and takeoff on the Moon. Although the LM had per­formed well on Apollo 9’s Earth orbital flight and Apollo 10’s close approach to the Moon’s surface, leaks in its propellant tank had only recently been fixed. With only two LM test flights under our belts, NASA management was still concerned about this problem. Our office was understandably chagrined at the

changes in the timeline and the science payload, but this turn of events lent even greater importance to ensuring that the science planned for the next landings was not compromised.

Another interesting exchange before a Senate committee took place shortly after the House subcommittee hearings. Homer Newell and John Naugle ap­peared before the Senate Committee on Aeronautical and Space Sciences on May 1. During the questioning, Senator Carl T. Curtis (R-Neb.) asked Newell and Naugle if knowledge gained from our completed space missions had changed previous beliefs. Both Newell and Naugle said yes, and Newell went on to provide a surprising example. He said that the “mascons” discovered by tracking the Lunar Orbiter flight paths (concentrations of high density material below the surface of the lunar seas that might indicate large meteor impacts) ‘‘give rise to some of the speculation that maybe at one time these areas were actually oceans or seas and [that] sediments from these oceans or seas is what filled those holes.’’ You won’t find these speculations in chapter 2, although many thought there was a chance that some water had been present on the Moon at one time. The theory that the Moon once had oceans was not sup­ported by any prominent theorists of the day, and if the large impact craters had been filled with sediment of some kind they would have been deficiencies of mass, not mass concentrations. For a crater to be a ‘‘mascon,’’ the fill had to be some unusually dense material. Even senior NASA managers had a hard time keeping up with changing theories as new information was gathered and ana­lyzed by more and more students of the Moon.

One week before Apollo 11 lifted off, Sam Phillips issued a new Apollo program directive (APD) detailing a total of ten lunar landing missions.2 The first landing was designated a G mission with the characteristics noted above, and the next four were called H missions. The H missions were designed around two EVAs, surface staytimes of up to thirty-two hours, and our old reliable payload of some 250 pounds. The final five missions, Apollo 16 through Apollo 20, were called J missions. Although the APD did not specify any science payload numbers, it stated that both the lunar module and the command and service module would be improved to permit longer staytimes. We anticipated that the LM would be able to carry additional descent propellant, which would translate in part to an ability to carry larger science payloads. We still held out hope in 1969 for flights beyond Apollo 20, but realistically we would have to extract as much science as possible from these ten missions. It was not exactly what we had planned for in 1964 and 1965, but we expected the J missions to be far better than the original Apollo plans. An interesting statement in the APD was that the constant-volume space suit would be available for the J missions. This never came to pass, and if such suits had been used they probably would have had little effect on the productivity of the J mission EVAs. LM and CSM consumables became the limiting factors, not the astronauts’ metabolic rates.

As we had simulated at Martin Marietta in 1964 and 1965 in case of an abort after touchdown, the crew of Apollo 11 first used their eyes to describe the lunar scene and took a few photographs before leaving the LM. One other piece of data collected was a movie of the landing site filmed from Aldrin’s window as Armstrong maneuvered for the landing. Not much scientific use was made of this movie because of its limited view of the surface, but you could see how the Moon’s surface layer was disturbed by the exhaust of the LM descent engine, with the fine-grained particles shooting rapidly away from below the LM in a fuzzy blur. These pictures confirmed that the lunar surface reacted as predicted to the LM exhaust and helped ease concerns about future LM landings. Peering out his small window, Armstrong provided the first descriptions of the surface, and Armstrong and Aldrin took pictures with the Hasselblad camera and de­scribed what they could see from their windows. Their observations added to the overall understanding of the landing site but did not reveal precisely where they had landed.

Whether Armstrong or Aldrin would have the honor of being the first human to stand on another celestial body had been decided long before Apollo 11 was launched. The initial timelines, circulated almost a year earlier, had indicated that Aldrin would be the first out. As planning for the mission ma­tured, however, it became evident that the LM commander, Armstrong, would be in the best position inside the LM to perform this historic first, seniority notwithstanding. From a science standpoint it really didn’t make any difference who would be first on the surface, but for Aldrin the decision was obviously a disappointment, and it continued to trouble him years later. Usually few people remember who was the second to do something; however, both Armstrong’s and Aldrin’s names are synonymous with the first Moon landing. Through the years Aldrin has received his deserved recognition, but he is not quite as famous as if he had been the first to touch the Moon.

It took some time for Armstrong and Aldrin to secure the LM and get it ready for a quick takeoff, if necessary. After landing and preparing for an emergency takeoff, the timeline scheduled a meal followed by a sleep period. The astro­nauts, understandably excited and not sure how long they would be permitted to stay on the Moon, asked Mission Control to skip the sleep period and immediately begin preparing for their EVA. Receiving approval, they donned their space suits, and a little under seven hours after they landed Aldrin opened the hatch. Armstrong squeezed through and bounced down the ladder (without seeing any exploding ‘‘Gold dust’’).

His descent and first steps on the Moon were recorded for all the world by a television camera attached to the landing stage, which he activated from the top of the ladder. This camera, built by the Westinghouse Aerospace Division, had been the subject of much debate. Could we afford the weight (about ten pounds) and the complications of deployment, since we knew the quality of the pictures would be poor? I was for not carrying it, especially when we were discussing whether to include the ALSEP because of weight and EVA time concerns. But once it was decided to eliminate the ALSEP, the question became moot from a science perspective. The ‘‘let’s carry it’’ side won the day, and it turned out to be a valuable tool both for public relations and for science. We used the TV pictures, in spite of their poor resolution, to help reconstruct the astronauts’ movements and plot the geology. Some senior NASA managers complained during the mission about the poor quality of the pictures, but by then it was too late. (The poor picture quality was caused not by any Westing- house design deficiencies but by the NASA specifications, dictated by weight and power constraints and antenna performance.)

After examining the LM and reporting its status, Armstrong began describ­ing the scene around him and his impressions of the lunar surface. Then he took a few photographs. He collected the contingency sample and put it in a pocket of his space suit, and he was soon joined by Aldrin to complete their carefully choreographed EVA timeline. At this point Armstrong removed the TV camera from the LM and set it up some sixty feet to the northwest, provid­ing a limited view of the landing site and of the astronauts’ movements as they went about their EVA. From this time until Armstrong and Aldrin reentered the LM, they performed all their tasks as planned. I won’t go into detail on what they accomplished; references listed in the bibliography describe these activities in great detail. Both astronauts performed all their scientific assignments better than expected under extraordinary conditions. One might think that the first

men to land on the Moon might not have their minds completely on the scientific tasks before them. One might expect them to be thinking about the upcoming liftoff, a maneuver never before attempted, which their survival depended on. Armstrong and Aldrin seemed to put such concerns out of their minds. They appeared to be completely absorbed in deploying the experiments, sampling, and describing what they were seeing and doing.

Aldrin placed the EASEP, the last-minute replacement for the ALSEP, on the surface about sixty-five feet south of the LM and in the same general area as the Laser Ranging Retro-Reflector. He had no trouble unfolding the solar panels and erecting the radio antenna, and once set up the experiment turned on automatically. Back on Earth, signals were received almost immediately, relayed to Houston from the NASA Manned Space Flight Network. We knew it was working because the seismometer recorded Aldrin’s footsteps as he walked nearby, but we hadn’t expected to receive so many signals.

The MSC and Bendix engineers manning the EASEP console in the Science Support Room (SSR) soon began to see a problem. The temperature of the seis­mometer package was rising faster than expected. It took some time to arrive at a probable cause, but they finally decided that dirt and dust were covering some of the surface, reducing its ability to reflect heat. Both Armstrong and Aldrin had commented on how far the soil would fly when they walked, as well as on how dirty their suits got during the EVA. While deploying the EASEP they had completely circled the experiments, so it was logical that some soil had coated the surfaces. Also, based on Aldrin’s comments, as we continued to track the rising temperature after their takeoff, it appeared he had placed the EASEP experiments closer to the LM than requested. We assumed that dust thrown up during the takeoff had also been deposited on the experiment surfaces. We kept our fingers crossed that the soil would not overheat the seismometer and had not obscured the small corner reflectors of the LRRR, making it difficult to bounce laser beams back to Earth.

These eventualities didn’t come to pass; the seismometer survived the rest of that lunar day (fourteen Earth days) and the following lunar night and came back on line for seven more days when the solar panels saw the sun again. The seismometer recorded several interesting events during its short lifetime, in­cluding the shocks of the astronauts’ backpacks hitting the lunar surface when they were thrown from the LM and the small ‘‘moonquake’’ when the ascent stage lifted off. Based on this performance, we could anticipate that the seis­mometers of the same design scheduled for the full ALSEP deployments would provide even more information during their much longer lifetimes.

In addition to still photographs, movies, and the Solar Wind Composition collector foil, a total of some forty-seven pounds of individual rocks, soil, drive – tube cores, and the contingency sample, all neatly packaged, finally found their way to MSC, where the staff at the Lunar Receiving Laboratory, and eventually the sample analysis principal investigators, eagerly awaited them. On the recov­ery aircraft carrier, the USS Hornet, the samples were divided into two batches and flown to Ellington Air Force Base in separate aircraft to ensure that some samples would survive in case one plane was lost at sea. There was always the chance we might not get back again to collect more samples. From Ellington, they were carried to the LRL.

The astronauts, wearing isolation garments that they donned in the CSM while awaiting recovery and transport to the Hornet’s deck, were immediately sequestered in a specially designed trailer lest they contaminate those around them with some deadly unknown virus. After the Hornet arrived at Hawaii, they too were flown back to MSC in their trailer along with two volunteer MSC doctors, to begin their one-month quarantine.

The samples, which had arrived before the astronauts, were carefully opened in the LRL, inventoried, and briefly described. In the meantime we were moni­toring the signals sent back by the passive seismic experiment and attempting to find the LRRR that the astronauts had left behind. This latter operation was not as easy as we expected, since the exact location of the landing site was not immediately known. Mike Collins had attempted unsuccessfully to locate the LM from orbit using the command module sextant. After analyzing the flight data and the returned photographs, we passed our best estimate to the LRRR PIs, and the LRRR was found on August 1, 1969, by the Lick Observatory in California.

On August 23, 1969, one month after Apollo 11 splashed down and the date when the astronauts were released from quarantine, George Mueller forwarded a memo to Clare Farley, James Webb’s executive officer, to be included in the report being sent to the president summarizing the results of man’s first foray to the Moon.3 In his memo, drafted in part by our office, he described the initial scientific results of Apollo 11 and summarized the program adjustments that would be made as a result of the mission. Included with the memo was a preliminary traverse map compiled by Gerry Schaber and Ray Batson of the United States Geological Survey using tapes from the lunar module’s television camera, photographs taken by the astronauts, and educated guesses based on what the astronauts reported from the Moon. The map sent to the White House had been further updated during the astronauts’ debriefings while they were still in quarantine. By this time photographs of the astronauts on the Moon and a few photographs of ‘‘Moon rocks’’ had circulated in all the newspapers and some magazines, so Mueller didn’t include any photographs of the astronauts with his memo, but he did include a photo of one of the returned samples. The Schaber-Batson map had just been completed and represented new informa­tion not yet made public, tying together everything the astronauts had done during their brief stay.

Short and to the point (five pages plus attachments), Mueller’s memo pro­vided an initial age dating of one sample (3.1 billion years), compared the chemical and mineralogical content of a few samples with that of the Earth, and offered the conclusion that the Earth and the Moon probably were formed ‘‘from the same whirling cloud’’ some 4.5 billion years ago. (It wasn’t clear where that comparison came from, but it wasn’t too bad a description if you agreed with the conclusion.) He also briefly discussed some results from the passive seismometer and LRRR; the latter experiment permitted the measure­ment of the Earth-Moon distance to within twelve feet as opposed to the best previous accuracy of about two thousand feet. (The accuracy of a few inches predicted in chapter 7 would come only after several years of ranging from three or more stations.) The last sentence we added to the memo was, we hoped, a thinly veiled plea to the White House to the keep missions going: ‘‘The indications thus far are that the Moon is a celestial body with complex structure, geology, and chemical history that may take considerable effort to unravel.’’

Mueller’s attachment summarizing planned program adjustments had an important effect on the subsequent missions. With the lunar landing mandate successfully completed, Mueller now proposed to slow the pace of the missions from one launch every two and a half months to one every four months. He stated that this not only would save money but would allow us to ‘‘increase mission flexibility and scientific return in later missions.’’ This was a welcome change to those of us planning the science and to the staffs at MSC and KSC, who had been working around the clock to support the shorter schedule. This would, we hoped, allow us to factor in some of the results of the previous missions while developing the objectives for each succeeding one and to alter the science payload and astronaut training accordingly. To a large degree we were able to do this on the last three J missions.

With the flight of Apollo 11 successfully concluded, General Phillips relin­quished his position as Apollo program director and returned to the Air Force. He was replaced by Rocco Petrone, who until this new assignment had been director of launch operations at KSC. Rocco, a West Point graduate, was a large man. He had been a backup tackle on two of Coach Red Blaik’s most famous Army football teams of the 1940s, which won thirty straight games before being defeated by Columbia in 1947, my freshman year. The teams featured ‘‘Doc’’ Blanchard charging up the middle or Glen Davis scampering around the end, at times behind the broad back of Rocco Petrone. He was listed in the game programs of the time as six feet one and 202 pounds; in the 1940s these were not intimidating numbers for a tackle, but he wasn’t exactly small. In 1969 he was a little more imposing, perhaps with a few more pounds than he carried in his playing days.

I don’t have many recollections of specific meetings with Sam Phillips, but I do remember calm, quiet, efficient status reviews that moved along quickly, with Phillips clearly in command—a management style much like George Mueller’s. Meetings with Rocco were different. He came to Washington with a reputation as a hard-nosed, hard-driving manager with his record at KSC—all Saturns launched successfully—a testimony to his management skills and his team’s ability. He had succeeded in what must have been a difficult environment under the early tutelage of the German-trained rocket scientists assembled by Wernher von Braun and KSC director Kurt Debus, both known to be sticklers for detail and perfect performance.

Rocco was the only senior manager I worked with who truly had a pho­tographic memory. If you gave him a ‘‘fact’’ related to your program during a briefing, woe unto you if you changed anything a week, month, or year later. Rocco would catch or challenge you, and he was almost always right. Rocco’s meetings were a little more lively than Phillips’s, especially if there were discus­sions of delays or unexpected changes. He was never shy about showing his displeasure, and it was reinforced by his imposing frame. Conference calls between Rocco and the NASA centers were always interesting. Usually they were arranged to discuss some critical problem, so by their very nature they were bound to be contentious. As we listened to Rocco asking questions in his distinctive high-pitched, singsong voice, we could visualize the speakers at the other end of the line squirming as they tried to justify some earlier position that he didn’t agree with. Rocco soon became our strong right arm and a defender of lunar science. Once he was convinced of the correctness of a scientific position, we seldom lost any ensuing argument with MSC. After Rocco’s arrival we really buckled down to expand and improve the science on the last three missions.

Flight readiness reviews (FRRs) were another area where Rocco ran a taut ship. Hosted by KSC, they were the final review, held about one week before a scheduled launch. Chaired by Chester ‘‘Chet’’ Lee, Rocco’s Apollo mission di­rector, they usually lasted one full day. There were representatives from all the NASA centers involved in the launch as well as the contractors and the required Department of Defense participants—a cast of hundreds. Every aspect of the mission from prelaunch preparation to splashdown and recovery was discussed in detail and checked off as being ready if it passed the rigorous review. Action items or deficiencies recorded during earlier mission reviews were carefully analyzed to be sure they had been properly attended to. This process might result in long debates, followed by documentation to prove problems had been resolved. Any items still open after the FRR were subject to a final review and structured sign-off before launch. Here is where Rocco’s photographic memory was put to the test. He would recall the smallest detail and ask penetrating questions. If the presenter could not answer to his satisfaction, someone had to leave the room and gather the missing information.

FRR attendance was carefully controlled. NASA senior management was seated at the front of the room, along with at least one of the astronauts who would be on the crew or serve as backup crew for the launch under review. Briefers with their supporters scurried in and out as called for by the agenda. For the J missions, I was entitled to a seat at the back of the room to take notes and perhaps pass on a discreet question for Chet Lee or Lee Scherer to ask. But the FRRs tended to be a one-man show, with Rocco calling the shots and the other senior managers like James McDivitt, Deke Slayton, and Al Shepard recognizing his mastery of the occasion. Everyone knew Rocco’s reputation for detail, and no facts or concerns were held back. We all understood that the lives of the astronauts seated in the room with us could be in jeopardy if the smallest problem went undetected or unsolved.

Hangar S became a kind of science headquarters at KSC as we approached the Apollo lunar missions. It was a little seedy looking on the outside—the paint was peeling and the large S was barely readable—but the inside was a high-tech workshop. As the name indicated, it was formerly a hangar at Cape Canaveral Air Station, but it now functioned as an important facility at KSC where final preparations and checks were carried out for all the experiments. Mock-ups of the LM and CSM were maintained in the hangar and used for stowage checks and simulations, which became increasingly complex for the missions follow­ing Apollo 11. The crews would spend more and more time at KSC as they neared the launch date, so it was important to have a place where they could stay up to date on any changes that might involve the experiments.

Flight experiments were sent to KSC from contractors around the country. KSC engineers would receive the flight hardware and store it in a clean room in another building near hangar S where final checks would be made to ensure that nothing had been damaged during shipping. Contractors building the experiments and equipment did their own inspections before the items left their plants, but the final checks were done at KSC. Nothing was loaded on the LM or CSM if it had not undergone a rigorous preflight inspection. Once it passed this inspection, it would be taken to the Vertical Assembly Building to be stowed.

Since ALSEP was the major science payload after the flight of Apollo 11, it received the most attention. It was carefully unpacked in the clean room, and each experiment was set up to check cable connections and any unique fas­teners, thermal blankets, or other apparatus that might give the astronauts trouble during lunar deployment. Chuck Weatherred, the Bendix ALSEP man­ager, recalled an important exchange as he helped the KSC team prepare for the launch of a ‘‘dummy’’ ALSEP on Apollo 10, scheduled to fly to the Moon but not land. Peter Conrad and Richard Gordon, the Apollo 12 crew, came into the clean room to watch the processing of the package that would simulate the weight and center of gravity of the ALSEP so that the MSC flight dynamacists could calculate how the spacecraft would react to various maneuvers during the mission. Although they had visited Bendix and seen their ALSEP in the final stages of manufacture, they knew their training schedules did not call for them to have any direct interaction with it until they were on the lunar surface. Conrad asked Chuck if they could participate in the final checkout before their ALSEP was stowed for the journey to the Moon. Chuck thought that was a great idea and said he would get approval from MSC, but Don Wiseman, his MSC contract manager, turned the request down. MSC didn’t want the astronauts fooling with the flight hardware before they deployed it on the Moon.

After several appeals and backing by the astronauts, that decision was re­versed, and all crews starting with Apollo 11 were permitted to work with the flight hardware at KSC before it was finally stowed for the trip to the Moon. It was perhaps a small victory, but I feel sure it made the crews more confident that they would not confront any surprises. ‘‘Murphy’s Law’’ says anything that can go wrong will go wrong. No matter how closely you monitor the manufac­ture of such a complex set of equipment as ALSEP, minor changes not reflected in the simulation hardware or documentation (someone’s last-minute bright idea) can creep into the design and could cause complications 238,000 miles away. We had few such problems with the science payloads, in part because we worked hard to be sure the astronauts were always in the loop.

At the same time that we were savoring the success of Apollo 11, the National Academy of Sciences’ Space Science Board was conducting another summer study, once again at Woods Hole. The study was chaired by Harry Hess of Princeton University, who had also led the 1965 summer study held in conjunc­tion with the Falmouth conference. Harry was a strong advocate of manned and unmanned lunar exploration, and his position at the Academy as well as his overall reputation in the scientific community lent great weight to our Apollo science planning. Harry’s objective for the study was to capitalize on Apollo 11 ’s success and lend support to those of us arguing with the administration and Congress to use the remaining Apollo hardware to carry out more missions and missions with ever increasing exploration potential.

Immediately after Apollo 11 ’s return, some leading decision makers in and out of Congress, who will remain unnamed, had been quick to propose ending lunar exploration and spending the money saved on various social programs back on Earth. These discordant voices motivated Hess to quickly call for the study. I attended the meeting with Don Wise, who had joined our office from Franklin and Marshall University to be Lee Scherer’s deputy. We made several presentations based on our ongoing efforts for the J missions, pointing out the potential for enhancing the science return. We also reviewed the recommenda­tions of the Santa Cruz conference and the ‘‘Lunar Exploration Plan’’ we had disseminated at the end of 1968. This summer study provided a new oppor­tunity to resurrect some of our old plans for long-duration missions that we were forced to abandon in 1968 for lack of interest by Congress and the admin­istration. Along with many other participants in the Apollo program, I strongly supported Harry’s views that we must make the case to take advantage of this opportunity—to squeeze as much science as possible from the Apollo program. After all, the major expenditures had already been made; using all the hard­ware, and doing it more efficiently, would entail adding only a small fraction to the total spent to date for the new science payloads and mission operations.

Tragedy struck the study on the first day, August 25, 1969. During the morning coffee break Harry complained of chest pains and left to see a doctor. He never returned. We were told he died peacefully at the doctor’s office. This, of course, spread a pall over our meeting. We had lost an irreplaceable leader whose vision had been, since the earliest days, a major force in our efforts to bring good science to the Apollo program. Only a few special people, including Ralph Baldwin, Harold Urey, and Gene Shoemaker, can lay claim to being fathers of lunar exploration, and Harry Hess belongs in that company. We continued our deliberations under a new chairman, Bill Rubey, the newly ap­pointed director of the Lunar Science Institute, and then issued our report.4 A case was made to support the launch of the nine missions still being planned at the time and to continue additional missions through 1975. The study con­cluded: ‘‘The decision concerning the nature of the lunar exploration program after the mid-1970s will hinge on the national commitment to manned space flight and on the significance of the scientific discoveries that emerge in the next few years.’’

While this report was in press, those of us advocating more Apollo science received another blow. Bill Hess resigned from his position as director of sci­ence and applications at MSC; he finally got tired of bucking the entrenched antiscience interests there. Tony Calio, who had earlier worked with us on Foster’s staff, took Hess’s place. When Tony left our office to go to MSC, we gave him a going-away party in Washington, wished him success in taking on such a difficult position, and looked forward to having someone at MSC who would be receptive to our interests. At the time, we didn’t know his appointment would adversely affect our relationship with MSC, but within weeks it became apparent. Tony quickly adopted the MSC line, and our relationship with MSC regressed to where it had been two years earlier. He became hard to reach by phone, and when we did get through he ignored most of our suggestions. He also developed an intense dislike for the staff at USGS. I never fully understood the reason for this antagonism—perhaps it was a holdover of earlier disputes between USGS and some of the staff he inherited. But this undermined USGS’s ability to support the upcoming missions for which members of the Field Geology Team had an ever increasing responsibility. It was only through their close relationship with the astronauts and others in the astronaut office that they were able to influence the geology content of the missions.

Returning to the remaining missions, Apollo 11’s success and a ringing endorsement from the National Academy of Sciences energized many in the science community to propose exciting new experiments for the remaining missions as we geared up to take advantage of a relaxation in some of the mission constraints. Until Apollo 11 returned safely, every Apollo engineer and system and subsystem manager was holding a little in reserve just in case it was needed. A little extra weight, a little extra available propulsion, a little extra performance margin. Slowly, with the help of the Bellcommers, these margins were identified and translated into increased science payload and more operat­ing flexibility.

The Schaber-Batson map was the first attempt, other than during simula­tions, to reconstruct in near real time what was happening on the Moon. Although during the Apollo 11 mission there was no direct exchange between scientists on Earth and the astronauts, based on our Flagstaff simulations we could see how this could be done effectively for the later missions. For Apollo 12 and the remaining four missions we tracked the astronauts in real time and had an up-to-the-minute map of their progress in the SSR. We coordinated our tracking with the flight controllers and medical staff monitoring the astronauts’ performance to ensure that their traverses would not overextend their life – support expendables. This monitoring was especially valuable during the last three missions, when the astronauts were often far from the LM and we had to be sure they had enough life support reserve to walk back if the lunar roving vehicle failed. For the science team it had another important aspect: it allowed us to relay suggestions for modifying the astronauts’ activities through the CapCom as they reported their findings and, at times, changed the timelines on their own initiative.

In September 1969 we advertised the opportunity to propose new experi­ments for the J missions that would utilize the LRV and the longer staytimes. This announcement, while directed primarily to missions 16 through 20, indi­cated that proposals to perform simple experiments on flights earlier than Apollo 16 would also be accepted.5 Perhaps the most ambitious aspect of this announcement was our optimism about where we would be permitted to target landing sites for the flights that would follow the initial landings. Scientifically exciting sites recommended by the Group for Lunar Exploration Planning (GLEP), such as the central peaks of Copernicus and the rim of Tycho, were included as candidates in the announcement so that proposers could consider their unique characteristics for their experiments.

With the arrival of Tony Calio and the immediate change (for the worse) in climate at MSC in regard to science, we began to lobby Rocco Petrone to push MSC to modify management’s responsibilities for science in the hopes that this would improve our working relationship. He talked to Jim McDivitt about making some changes. At the end of October 1969 our office originated a memo for Petrone’s signature formalizing these suggestions. The opening sen­tence, underlined, stated, ‘‘I think we have a problem in the management of the science program which warrants immediate action.”6

McDivitt responded two weeks later and gave us half a loaf.7 He moved the design, development, testing, and delivery of approved Apollo experiments from Calio’s office, the Science and Applications Directorate, to the Engineer­ing and Development Directorate, managed by Max Faget. (We weren’t sure if this was a victory.) S&AD would still be in charge of the scientific requirements, science mission operations, postflight data analysis, and interactions with the PIs, but McDivitt promised that his office would strengthen its science over­sight. This was encouraging, since Petrone and McDivitt usually agreed on the important aspects of the missions, and science would take center stage for the remaining flights. In spite of these changes, our concerns would soon be echoed by the scientific community.

Through 1970, we were still hoping dual-launch missions might be rein­stated, enabling fourteen-day stays on the Moon, and the trade journals of the day continued to write about these plans as if they were approved.8 In Lee Scherer’s office we continued to study an LM shelter and a dual-mode (manned and automated) roving vehicle. Scherer urged Marshall Space Flight Center to complete the preliminary design and promised funding for this work.9 Mean­while, preparations continued for the next landing. Apollo 12, we hoped, would allow us to accomplish some of the science originally scheduled for Apollo 11 but at a different mare site, many miles to the west.

Apollo 12 was successfully launched in November 1969 and landed about eight hundred miles west of Tranquility Base at the lunar feature called the

Ocean of Storms, another mare site. If our photo interpretations were correct and the landing site was on an ejecta ray from the crater Copernicus, a few hundred miles to the north, we hoped to return samples of material from deep within the Moon, excavated by the impact that formed this huge crater, some forty miles in diameter. Copernicus is one of the craters you could identify under proper lighting conditions with your ten-power binoculars, just a little west-northwest of the center of the Moon.

Two EVAs were scheduled and carried out, and a full ALSEP was deployed. Peter Conrad and Alan Bean proved to be enthusiastic lunar explorers. Much was made in the press of Pete’s laughing, giggling, ‘‘cackling,’’ and joking as he went about his tasks, but he and Al performed flawlessly, bringing back some stunning pictures and a wide assortment of lunar rocks. The TV camera, simi­lar to the one carried on Apollo 11, was damaged soon after they climbed down from the LM, so we were completely dependent on their oral descriptions to reconstruct where they were and what they were doing. Our simulations at Flagstaff and at other locations once again paid off, and we produced a map of the landing site in the SSR based on their descriptions and dead reckoning of how far they traveled between sampling stations.

In addition to the sample collecting, a major objective of Apollo 12 was to land near enough to Surveyor 3 to allow the crew to walk to it and take pictures of the landing site for comparison with the Surveyor TV camera pictures sent back to Earth two and a half years earlier. They would try to bring back pieces of the spacecraft, including the TV camera mirror and scoop, so we could study the effects of thirty months of exposure to the lunar environment. The trajec­tory engineers in mission control and Pete’s piloting skills put the LM right on target, within a few hundred feet of Surveyor 3. This demonstration of the ability to land at a precise point on the Moon, as opposed to Apollo 11’s overshooting the landing point, eased some of management’s concerns as we advocated more difficult future sites. All objectives of the mission were met, and the ALSEP became the first link in the network that the geophysicists had dreamed of for over five years. By the end of their two EVAs, Conrad and Bean had successfully deployed the ALSEP (they encountered a minor difficulty while removing the fuel cask of the radioisotope thermoelectric generator from its stowage on the LM, but deployment proceeded as planned), retrieved pieces from Surveyor 3, and collected a wide variety of samples totaling some seventy – five pounds.

While Conrad and Bean were on the lunar surface, Dick Gordon, the CM pilot, was carrying out his tasks. Soon after the others landed he used his sextant to search for the LM on the surface and was successful, even observing the much smaller Surveyor 3 a short distance away. His primary job was to photograph the Moon from orbit using a Hasselblad and a new camera array called the Multispectral Photography Experiment. The array consisted of four 70 mm Hasselblad cameras with fixed focus, each equipped with a different filter to return photographs in the blue, red, green, and infrared portions of the optical spectrum. This camera array was flown originally on Apollo 9 with Paul Lowman as PI. (For Apollo 12, Alex Goetz of Bellcomm was PI.) Gordon would point the array through one of the CM windows and trigger all four cameras simultaneously every twenty seconds. The major objective was to photograph potential landing sites and, we hoped, use the pictures to extrapolate the re­turned samples to wide areas of the Moon based on spectral differences caused by compositional variation in the lunar soil and rocks. A good concept, but the Moon was not cooperative. When the photographs were developed subtle dif­ferences between the crater Lalande and Mare Nubium were found at only two points. We would have to wait until the J missions, with their more sophisti­cated sensors, to have this exploration technique pay off.

During debriefings of the Apollo 12 crew we asked why they had moved some of the rocks they sampled before documenting their location with photo­graphs, the preferred technique. Their answer was simple and logical. During their early sampling, they had found that many of the rocks they picked up and had documented were too large to fit into the sample bags. Because they were half buried their full size could not be estimated—they were like ‘‘the tip of an iceberg.’’ Rather than waste time photographing samples they could not save, they elected to lift some of the rocks before taking the requested six photo­graphs. As a result of this crew observation, the photo documentation require­ment for the next mission, Apollo 13, was reduced to five per documented sample (although that crew never had the opportunity to use the new standard) and continued to be revised, downward for subsequent missions as we better understood the documentation needs for mapping and cataloging the samples in the LRL.10

Although Gene Shoemaker was still officially the PI for field geology on this mission, Gordon Swann took over crew training and led the interaction of the Field Geology Team with the crew. (Swann would later be named PI for Apollo 14 and Apollo 15.) We exercised the crew at the Cinder Lake Crater Field simulation site outside Flagstaff, described in chapter 9, and other sites, and by mission time Swann and his team had established a good relationship with Pete and Al. They had both been good students, and their training carried over to the lunar surface. In addition to Pete’s enthusiastic, nontechnical descriptions of what he saw, he and Al also provided a good specific commentary that we easily followed, and the Field Geology Team was able to construct a real-time geologic map of the landing site.

After the mission returned we received a letter from a research physicist at the Atomic Energy Commission’s Lawrence Radiation Laboratory in California highly critical of the astronauts’ oral descriptions and their apparently poor training. We always responded to letters from the public on any subject. I was assigned to write a letter back for Rocco’s signature, and it seemed clear to me that the criticism was based on the press reports of Conrad’s voice transmis­sions, not on the whole transcript.11 In the response I included some of the astronauts’ descriptions not carried by the press, such as the characterization as “granitelike” the various colors they reported, and many other precise descrip­tions of rock shapes and soil conditions on the lunar surface. I hoped our response was reassuring to this concerned taxpayer. It was meant not to belittle his concerns but to show that this aspect of the missions—the astronauts’ geological training—was being seriously pursued so that based on their obser­vations we could extract a vast amount of information from each mission.

Apollo 12 had already gone to the Moon and returned before we were pre­sented with the detailed analyses of the Apollo 11 lunar samples. This delay was dictated by the quarantine requirements and by an agreement with the sample PIs not to release their findings until a formal conference could be held in January 1970, when all the results would be available.

Two months later, in March 1970, a new solicitation was issued that required scientists wishing to analyze lunar samples to submit, or resubmit, proposals to receive samples returned by Apollo 14 and subsequent missions. John Pomeroy joined our office at this time to manage the expanded sample analysis program and oversee the operation of the Lunar Receiving Laboratory. By July we had received 383 proposals, including proposals from 175 of the 193 teams (the number had grown from 142) that had analyzed samples from Apollo 11 and Apollo 12. Foreign interest in doing analyses was also growing, and of the 208 new proposals, 95 were from foreign investigators. Gerald ‘‘Jerry’’ Wasserburg, a sample PI from Caltech, writing to administrator Tom Paine in June about his recent trip to Europe, reported that ‘‘there is a fantastic amount of enthusiasm by all the scientists who are involved in these different countries, and. . . the foreign press has given them a tremendous amount of coverage. Some individ­uals, in fact, have become sort of national heroes.’’12 As before, almost all the proposals received were accepted, and many of these investigators and their successors still attend the annual conferences at the Lunar and Planetary In­stitute in Houston.

Before any of the missions, toward the end of 1964 I proposed to NASA management that we study the possibility of commanding the discarded LEM ascent stage to strike the Moon near seismometers that would be placed on the lunar surface by future astronauts.13 At the time, there was no plan to control the impact point of the ascent stage; if not controlled, it would gradually lose altitude and hit the Moon at some unknown time and place. If we could control the impacts of the LEMs, we would have the equivalent of large explosions that would be recorded by the network of seismometers we hoped would soon be in place. We could not be sure when a moonquake or a meteor might provide an energy source large enough to let us study the Moon’s interior. The seismome­ter packages would have finite lifetimes to record some large natural event; if such events were rare, and if the seismometers malfunctioned, they might not be operating when one occurred. Also, the ascent stage was a rather flimsy, lightweight structure, and I feared its impact might not be recorded if its natu­ral decay from lunar orbit occurred some distance away or perhaps even on the Moon’s farside.

We began to explore this idea with MSC and enlisted the support of Frank Press, Bob Kovach, and Maurice Ewing, all members of the seismic teams. It took several years to obtain approval for this maneuver, but by the time Apollo 12 flew we had an agreement to control the impact point of the ascent stage by using the fuel remaining after rendezvous to make it leave orbit at a planned point. For Apollo 12 we recorded the astronauts’ movements and LM takeoff on the ALSEP seismometer as we had for Apollo 11, after which the Moon settled down again and was quiet until the ascent stage hit five hours later, about forty miles away.

We calculated that the impact was equal to setting off an explosive charge with an energy equivalent of about one ton of TNT. The first seismic wave arrived at the Apollo 12 ALSEP 23.5 seconds after impact, building to a maxi­mum amplitude about seven minutes later, with the total recorded event lasting some fifty minutes. The signal recorded was unlike any seismometer recording observed on Earth after either a manmade or a natural event, especially if one considered the relatively small amount of energy involved. This led to a number of theories about the unusual composition of the Moon’s outer layers that might cause such a response. We would have to wait for more information, gathered by the next ALSEPs, before a model of the Moon’s interior finally emerged that most geophysicists would agree with. When we described to George Mueller the effect of the LM impact and the unusual response, he said, tongue in cheek, that the large amount of titanium found in the lunar samples suggested the Moon must be a hollow titanium shell—a spacecraft from an­other galaxy covered with cosmic flotsam and jetsam.

At the end of 1969 Mueller resigned. He had steered the Office of Manned Space Flight, and NASA, to its improbable goal of landing men on the Moon and bringing them safely back to Earth. His management skills have been described by many, and I hope I have given a few insights that will add to an appreciation of those skills. Like Rocco Petrone, he embraced the importance of ensuring that good science be accomplished on the missions. Although I have never been able to ask him why he left NASA, I would not be surprised if a major reason was his frustration at failing to persuade the political powers to approve a long-range plan for continuing manned exploration to the Moon and Mars using the capabilities he and many others had worked so hard to build.

He was replaced by Dale Myers, who had been North American Rockwell’s manager for its Apollo spacecraft contract. Dale had survived both the bad times at Rockwell, when the contract was in trouble for many reasons, and the good times starting with the success of Apollo 8. It must have been a major culture shock to move from being a contractor who had to bow to his NASA ‘‘bosses’’ to being in charge. But he handled it well, and he had a seasoned team to lean on in his first days. I participated in a number of briefings for him early on, and we hardly skipped a beat as we brought him up to speed on all aspects of the program. He selected Charles Mathews as his principal deputy and Charles Donlan as his technical deputy; both were old NASA hands who could help him understand some of the pitfalls he faced. Eight years later, after we had both left NASA, our professional paths would cross again when Dale was appointed undersecretary of the newly created Department of Energy and I was his acting assistant secretary for conservation and solar energy.

Myers’s appointment was only one of several major senior management changes made at this time. Other new blood included George Low, whom Tom Paine, Webb’s successor, brought to Washington from MSC to be deputy ad­ministrator. All these changes had little effect on the upcoming flights. It did seem, however, that once in Washington Low became more sympathetic to the needs of the scientific community, and he strongly supported the efforts to place a high priority on the scientific returns from the final missions.

Once we had an agreement to control the impact of the LM ascent stage, after Apollo 11’s successful mission I proposed deliberately targeting the upper stage (the SIVB) for a lunar impact. This was a lot harder sell than controlling the impact of the LM ascent stage. The SIVB stage, as described in chapter 5, was programmed to deliberately miss the Moon. If it was maneuvered for an impact after placing the CSM and LM on a translunar coasting trajectory, it would arrive at the Moon about the same time the astronauts would be braking into lunar orbit. This was why the original mission rules called for the SIVB to be steered away from the Moon after translunar injection, to avoid any chance that it might interfere with the CSM and LM.

Asking that these rules be changed raised several safety concerns. Not only would the CSM with attached LM and the SIVB be traveling near each other toward the Moon, but it was feared that the powerful impact of the SIVB might hurl debris high above the Moon into the path of the CSM and LM. We asked MSFC to determine if sufficient propulsion would remain after translunar injection so that we could steer the stage and if there would be any problems sending commands to control its trajectory. Douglas Aircraft Company, the SIVB manufacturer, had studied such an application of the Surveyor translunar insertion stage when it was thought that the Surveyor spacecraft would carry seismometers to the Moon, so some of the homework had already been done.

MSFC came back quickly with an analysis that it could be accomplished; it was only too glad to have this opportunity to demonstrate its engineering prowess and the versatility of one of its babies. At the end of May 1969 MSFC made a presentation to me and Michael Yates, and at the end of June we presented our case to the Change Control Board, providing the analyses show­ing that the SIVB could easily be commanded to hit at a preselected point and that debris from the impact would not threaten the LM and CSM.14 Approval was given to proceed with the SIVB modifications, to the delight of the passive seismometer team. We would have to wait until Apollo 13, scheduled for an

April 1970 launch, before all the changes could be made to the SIVB and its command software to achieve the controlled impact.

After Apollo 12, the ‘‘rump GLEP’’ and GLEP came into conflict with the conservative MSC engineers. Some of the sites on our list for the remaining eight missions would require maximum performance from all the Apollo com­ponents. I can recall a contentious meeting at MSC, shortly after Apollo 12’s return, when the subject of future landing sites was on the agenda. This was a meeting of MSC managers and engineers to which a few of us from headquar­ters and Bellcomm were invited. Bob Gilruth, MSC center director, was the senior manager present, but the meeting was run, as usual, by Chris Kraft, Gilruth’s newly appointed deputy, and by Jim McDivitt, manager of the Apollo Spacecraft Program Office. Jim, a recently retired astronaut, was an excellent manager and ran a tight ship. Among other qualities, he was noted for his famous (or infamous, depending on your point of view) daily status reviews, held in a conference room lined with displays and charts and devoid of chairs: no nonsense, get the information out, assign actions, and get back to work! The only bow to comfort was a long table down the middle of the room where you could set your coffee cup while you took notes. Based on his positive response to Petrone’s earlier letter, we considered Jim relatively neutral in our debates on how to accomplish the best science. His major concern was always crew safety; if safety was not compromised, he would usually support our requests.

After the near pinpoint landing of Apollo 12, some of the constraints on site selection described in chapter 5 were relaxed, in particular the requirement for multiple sites to accommodate possible launch aborts. Only one backup site had been designated for Apollo 12, about thirteen degrees farther west, which would have allowed for a one-day recycle if a problem had occurred before launch. The rump GLEP and GLEP went through a process similar to our earlier deliberations to select high priority sites for landings after a successful Apollo 12. This time we came up with a new set A including seventy-two sites. We then narrowed the list to a set B of twenty-one sites and finally recom­mended twelve that included Fra Mauro for Apollo 13 and even more challeng­ing sites for missions 14 through 19. (By now the number of landing missions had been reduced by one, but we were still planning on a total of nine landings.)

But back to the meeting. Equatorial sites had been agreed on for the first three landings as the safest and most easily accessible, although the Apollo 13 site, Fra Mauro, would be a little more challenging since it was surrounded by

rougher terrain. These initial sites were within the “Apollo zone of interest.’’ All were close to the equator and were covered by the greatest number of high resolution Lunar Orbiter photographs. Many uncertainties still existed in pre­dicting the performance of the total Apollo system, but Bellcomm had already completed an analysis of SIVB, LM, and CSM performance showing that a high percentage of the Moon’s nearside could be reached while maintaining the required safety margins.

As the meeting droned on and such things as communication restrictions and propulsion budgets and margins were discussed, it became apparent that MSC management was going to take a conservative stand. Those of us who had been working on future landing sites were being asked (not quite directed) to rein in our expectations and continue to look for science sites near the Moon’s equator. The nearer to the equator you landed, the more options were available to get you out of trouble. There was reluctance to go outside the ‘‘Apollo zone’’ despite the Bellcomm study. Besides, it was a Bellcomm analysis, not one done by MSC engineers.

MSC’s position was certainly understandable. Every mission was risky, from liftoff to splashdown, and a difficult lunar landing site only added to the risk. No one wanted to be responsible for the decision to land at a site where a crew would be lost, for whatever reason. An accident, such as befell the crew of Apollo 1, could result in the cancellation of the remaining missions, an outcome that few in NASA would have cheered. For the staff at MSC each flight involved more personal worries than, perhaps, for someone in Washington or elsewhere in the scientific community; crew members were their neighbors and co­workers. If a crew didn’t return they would be living with the grieving families.

By this time I had many close friends in the astronaut corps and fully appreciated the danger inherent in each mission. However, Noel Hinners and I felt obliged to speak up. The only rationale for continuing the missions was to carry out good science, and this could be done only if we were allowed to explore sites far from the equator, sites already identified as having the potential to resolve important questions. We went so far as to predict that, based on Lunar Orbiter photographs, safe LM landing sites could be found almost any­where on the Moon. If any other proof was needed, look at Surveyor 7, which, with minimum ability to target the landing site, had managed to land in rough terrain on Tycho’s rim without an astronaut making last-minute adjustments. How much easier it should be with a man at the controls. These remarks were met with skepticism and grumbling from around the table, but this position was gaining support from many others, including some of the astronauts.

Eventually, as others with more clout weighed in, MSC management reluc­tantly agreed to process sites away from the equator. Undoubtedly each mission that lifted off after Apollo 13’s near disaster increased their anxiety; the chances of a major problem were rising with each flight. No matter how carefully we prepared, one or more of the five million parts included in every launch vehicle and spacecraft could fail or malfunction at any point in a mission.

On March 6, 1970, the Apollo Site Selection Board met at KSC to select the landing site for Apollo 14. With Apollo 13 scheduled to land in the western part of the ‘‘Apollo zone,’’ this was the first meeting of the board since the meeting described above. We looked on it as a test to see if MSC management would be swayed by our arguments and allow Apollo 14 to land outside the ‘‘zone.’’ Tony Calio, who had replaced Bill Hess as chairman of GLEP, presented the results of the GLEP meeting of February 6 and 7. GLEP recommended a site called Littrow, at the southeastern edge of Mare Serenitatis, well north of the ‘‘Apollo zone,’’ and the MSC in-house site evaluation team recommended the same site. After several presentations, including two by Lee Scherer and Noel Hinners, the board approved the Littrow landing site, and Jim McDivitt signed off in agree­ment.15 We had overcome the last hurdle toward planning the scientific explo­ration of the Moon during Apollo.

A key science ally at MSC was Jack Sevier. Jack’s personality was perfect for the difficult job he was assigned, acting as a mediator between the scientists and MSC’s engineers. Easygoing, with a ready smile and quiet sense of humor, Jack had been an important contributor to the rump GLEP meetings starting in 1967, providing MSC’s views on the constraints that could affect site selection. He was the branch chief of the Operations Analysis Branch and as such was the focal point for all the competing factors that could influence the outcome of our scientific activities. He would later lead the Lunar Surface Planning Team for the J missions, which developed the astronauts’ lunar surface timelines and ultimately shaped the successful outcome of each EVA.

With the missions still being scheduled rather rapidly and changes in their scientific content occurring with each mission, some members of the scientific community continued to publicly criticize how Apollo science was progressing. Soon after the return of Apollo 11, Gene Shoemaker was quoted as being highly critical of the way NASA management treated science on the Apollo missions.

This view troubled me deeply at the time: we had been working hard to expand the science, and I knew he was aware of how much more productive the next missions would be. There is no question they could have been better, but we had made great progress since he had first become involved. His statement drew the ire of Homer Newell and Rocco Petrone. Harold Urey, perhaps egged on by Tommy Gold, who always seemed to delight in knocking NASA, also criticized the lack of scientific input into NASA decision making.

In a letter to Newell in March 1970, Urey said he agreed with Gold ‘‘that well known people who have been concerned about the moon for years are so systematically neglected by the management of NASA.’’16 He was particularly irate at their exclusion from the selection of Apollo landing sites. In regard to site selection he wrote that ‘‘the people who vote are loaded with geologists of a very limited view of lunar science,’’ and he made a few other scathing com­ments. By this time, after just two missions, Urey was seeing the writing on the wall. His well-publicized theories on the Moon’s origin were being proved wrong, and I suppose Nobel laureates don’t like to be proved wrong. At a later date he would acknowledge his errors and even make jokes about them.

Newell’s staff was asked to respond to Urey’s letter, but they sent an informa­tion copy to our office. Rocco Petrone, not taking kindly to this criticism, asked that we address one of Urey’s comments dealing with site selection. In my memo for Petrone’s signature, which we hoped would be included in Newell’s formal response, I named the scientists and engineers present and voting at the last Site Selection Subcommittee meeting at MSC.17 I listed twenty-one names: three geologists, two astronomers, four geophysicists, three NASA engineers, two geochemists, one nuclear chemist, three physicists, one geodesist, one atmospheric physicist, and one cosmologist-chemist, Harold Urey. Of the twenty-one, nine were government employees or contractors and the other twelve came from universities or private research laboratories. All had been involved in lunar research for at least the past five to ten years, which pretty well covered the period when interest in the Moon became widespread. Urey had picked the wrong topic—site selection—to complain about, but his overall concern had some merit. His complaints and those of others were primarily a criticism of how MSC was interacting with the scientific community, which once again was becoming contentious after Tony Calio replaced Bill Hess.

Urey’s letter came just one month after a meeting at MSC when a group of

scientists, all closely involved in Apollo investigations, met with MSC manage­ment to discuss the problems they were having working with MSC staff. Urey had not been invited to this meeting, nor had Tommy Gold, which may have added to their pique; Newell attended as an observer. After the meeting Newell apparently thought the situation was resolved and wrote Gilruth a complimen­tary letter; but he didn’t really understand the depth of distrust that was build­ing between Calio’s organization and the scientists who were devoting more and more of their time to making each mission as successful as possible. Yet the meeting was useful in making McDivitt and Chris Kraft more aware of the needs of the scientists, and relationships with their offices improved. MSC agreed to arrange for more time in the astronauts’ schedules so the PIs could explain their experiments and their requirements during deployment or opera­tion. The PIs also asked for a better system of communication between the scientists in the SSR and the crews. They cited difficulties that arose during Apollo 12, when it took ten to fifteen minutes for questions raised in the SSR to be relayed to the astronauts by the CapCom, if they went out at all—and often they didn’t.18 There was some improvement on succeeding missions, but in general MSC and the Flight Operations Directorate (FOD) tended to ignore this latter request. Flight directors and CapComs felt, with some justification, that they shouldn’t interrupt the crews on the surface with a lot of questions and directions; they had enough to think about.

Apollo 13 was a scientific disappointment but an engineering triumph. We lost a precious ALSEP (one of only six purchased), but the opportunity to study this site and collect valuable samples was realized when Apollo 14 went back to the Apollo 13 landing site. In spite of this disappointment, I never heard any complaints from the PIs, many of whom had worked with Jim Lovell, Fred Haise, and John ‘‘Jack’’ Swigert to prepare them for their flight. Like everyone else, we could only cheer the skill of all the NASA engineers and support contractors who brought the crew home safely. The Apollo 13 crew members who performed so well under the threat of being the first astronauts to die somewhere in space, and the many heroes in the FOD led by Eugene Kranz, have had their roles well documented, so I will not try to add to that story. Science probably gained from the failed landing. It helped us refocus on how important each mission was. There were no givens; we had to make sure the remaining missions would be fruitful. And it seemed to make management more receptive to our requests to improve the science content of the last mis­sions. The drama of Apollo 13’s rescue also ensured a more attentive public for the next missions and a wider audience interested in what we were discovering.

One experiment, the passive seismometer left behind at the Apollo 12 land­ing site, did achieve important results from Apollo 13. Despite the problems the crew encountered during the rest of the mission, the Apollo 13 SIVB stage, the first programmed to strike the Moon, accomplished its job by landing some eighty-five miles from the Apollo 12 ALSEP. The seismometer received strong signals, and the impact had so much energy-estimated to be the equivalent of twelve tons of TNT (larger than the LM ascent stage impact because of its greater mass and higher velocity at impact)-that it sent seismic waves deep into the lunar crust. This elated Gary Latham, the passive seismometer PI, because he and his coinvestigators could now make some preliminary estimates about the Moon’s deep structure.

When Lee Scherer’s office was formed at the end of 1967, several of us involved in lunar science planning left Advanced Manned Missions, but Phil Culbertson stayed, eventually becoming director. In March 1970 he negotiated a memorandum of understanding with the Apollo Program Office to work cooperatively on lunar planning in case funding became available to continue missions beyond the scheduled Apollo flights.19 Our two offices continued working jointly on post-Apollo planning for several more years, despite the lack of official sanctions to build the hardware needed for extended missions.

After Apollo 13 failed to land, and reacting to the increasing clamor in some circles to halt the missions, in July 1970 our office issued a summary report of what we had learned to date from all our missions, manned and unmanned, and where we thought lunar exploration should be going.20 The objective of the report was to support Culbertson’s planning efforts in Advanced Manned Mis­sions and to present an “Integrated Space Program Plan’’ that would provide mission schedules extending to 1990. It represented our last effort to justify a continuing program of manned and unmanned exploration by building on Apollo and other programs, including Mariner and Viking, and factoring in programs on the drawing boards such as Skylab and space stations. We pre­sented an integrated program that included lunar bases and manned inter­planetary launches.

Recently a quotation from Charles Lindbergh came to my attention. Asked about the $25,000 Orteig Prize offered for the first nonstop flight between New

York and Paris, which he won with his daring flight in 1927, he responded, ‘‘The important thing is to start: to lay a plan, and then follow it step by step, no matter how small or large each one by itself may seem.’’ One could make a reasonable argument that Lindbergh’s successful flight was the first step toward today’s commonplace travel across the Atlantic and to almost every point on the globe. With Project Apollo we had taken the first step in mankind’s leaving Earth and exploring our solar system. We believed we had put forth a step-by­step program to build on Apollo and move logically to the next objectives: space stations, lunar bases, and manned flights to Mars as early as 1989.

No such logical plan was ever agreed to. Some administrations have ignored space exploration, and some have paid it lip service. In the end, a program that would take advantage of the expertise and capabilities developed for Apollo was never endorsed. The report is now resting in one of my dilapidated packing boxes, perhaps the only surviving copy of our vision of a long-range plan for exploring the solar system. It was grandiose—undoubtedly too grandiose for the times—but in 1970 everything we proposed was achievable based on the technology in hand. All that was needed was the leadership to commit the nation to the next step.

In January 1971, just two weeks before the scheduled liftoff of Apollo 14, the second lunar science conference was held at the Lunar Science Institute. Al­though many of the same people attended this conference as were at the one held after the study of the Apollo 11 samples, the sense of excitement was missing. The only new samples that had been studied, aside from a few grams of material brought back by the Soviets’ Luna 16, were those returned by Apollo 12 a year earlier. Whereas restrictions had been placed on the release of infor­mation about the Apollo 11 samples, the Apollo 12 sample PIs were not pre­vented from publishing the results of their studies of material returned by the mission. Most of the new information was already public and well known by the attendees.

The big debate at the conference dealt with the significance of the high content of radioactive elements (uranium, thorium, and potassium 40) found in some of the Apollo 12 samples, which would imply an early, very ‘‘volcanic’’ Moon. There were other differences from the Apollo 11 samples, suggesting that the Moon may have had an unusual differentiation history. It also appeared after initial study that the mare material sampled at the Apollo 12 site was about a billion years younger than that collected at the Apollo 11 landing site, suggest­ing that the Moon had gone through several major periods of mare formation. These findings would continue to be debated as each mission brought back new information.

Apollo 14 was sent to the site chosen for Apollo 13, Fra Mauro, in a hilly, upland area just a short distance east (112 miles) of the Apollo 12 site. From the perspective of our plans to deploy the ALSEPs in a broad network so we could triangulate on phenomena at the Moon’s surface or occurring at depth, being so near the Apollo 12 ALSEP was not ideal. But from a geological point of view it was considered an important site, since we believed that the samples returned would include debris ejected from the huge Imbrium basin to the north. Again, as for Apollo 12, we hoped to collect samples from deep within the Moon that would help resolve some of the questions raised at the second lunar science conference. They would also be useful to Gary Latham and his coinvestigators in interpreting the Moon’s deep structure, since these rocks would tell them how fast the seismic waves created by the SIVB impacts should travel compared with what they were observing in the records received back on Earth.

Had Apollo 13 been successful, we were willing to accept the deployment of the ALSEP so close to Apollo 12. It was to be the last of the landings near the Moon’s equator, reflecting MSC’s cautious approach. We had expected that after Apollo 13, Apollo 14 would land at Littrow, the first site selected solely for its scientific value and, because it was far off the lunar equator, ideal for our ALSEP network. The geological rationale for landing at Fra Mauro still held, but the decision to retarget Apollo 14 there was doubly painful from a scientific perspective. With the loss of Apollo 13, there were only six more projected landings (ultimately reduced to three) to uncover the Moon’s secrets hidden on or below a surface area roughly equivalent to all of North and South America combined. And well over half of that area was inaccessible because it was outside our landing capabilities or on the Moon’s farside. Imagine trying to understand those two continents with only six widely scattered small points of knowledge!

The landing site was to be within walking distance of what appeared to be a crater of recent vintage, named Cone by the Field Geology Team because of its steep, funnel-like inner slopes. From the Lunar Orbiter photos we could see large blocks on the rim of Cone Crater, reinforcing the belief that if the astro­nauts could get to the rim they would be able to sample Imbrium ejecta in the rocks ‘‘mined’’ by the Cone Crater impact. Alan Shepard guided the LM to a perfect landing within two hundred feet of the target point and less than a mile from Cone Crater, whose rim could be seen in the distance when he and Edgar Mitchell descended the LM’s ladder. This time there was a color television camera, with better resolution than the Apollo 11 camera, and it functioned well, providing views of the astronauts as they climbed down to the surface and panoramas of the landing site as they worked near the LM.

Between the Apollo 13 and Apollo 14 launches we had built a small two­wheeled cart, the modularized equipment transporter (MET) discussed in chapter 8, to help the astronauts carry all the gear that was now part of the field geology experiment. It was unloaded from the LM descent stage near the begin­ning of the first EVA, and the crew stowed the tools and equipment they would need for the sampling scheduled on the first EVA and the traverse to the rim of Cone Crater, the major objective of the second EVA.

The first EVA went off with no big hitches, and the major tasks—the ALSEP deployment and sample collection near the landing site—were successfully completed. A new experiment, the active seismic experiment, was conducted in conjunction with the ALSEP deployment. Three geophones were strung out on cables to the south of the ALSEP, with the last one approximately three hundred feet from the ALSEP central station. The first part of the experiment consisted of setting off small charges, about the size of a shotgun shell, housed in a hand­held ‘‘thumper’’ hardwired to the ALSEP central station electronics, which provided timing data and transmitted the signals received by the geophones back to Earth. Mitchell carried the thumper out to the last geophone and set off a charge, then retraced his steps back to the geophone closest to the central station, setting off charges along the way. Twenty-one charges were scheduled, but a few misfired and only thirteen were recorded. A second part of the experiment consisted of a mortar designed to fire four small explosive charges various distances away from the geophones, the farthest to land five thousand feet from the mortar. This second part of the experiment was not conducted until many months later, to avoid any possibility that the mortar fire might damage the nearby ALSEP central station.

Although Shepard and Mitchell could see the ridge formed by Cone Crater in the distance when they started out on the second EVA, once they began walking and pulling the MET they soon lost sight of the ridge behind the intervening low hills and hummocks. Others have described in some detail their difficulties in reaching the rim of Cone Crater. They didn’t quite make it, but they came close, and they sampled ejecta thrown out by the impact that formed the crater, the main geological objective of the mission. After the diffi­culties they encountered attempting to reach Cone Crater’s rim, they probably both wished they had the LRV that would be carried on the next mission. Another new experiment on this mission, the Lunar Portable Magnetometer, was operated twice during this EVA, and the readings were relayed back to Houston by voice. The samples collected during both EVAs weighed almost ninety-five pounds.

Like the CM pilots before him, Stuart Roosa carried out several experiments on the way to the Moon and while the other astronauts were on the lunar surface. The number of experiments assigned to the CM pilot was increasing with each mission as we attempted to take full advantage of his time and the added payload weight that was becoming available. Roosa completed several new photographic tasks and other types of experiments. Bellcommers Farouk El Baz and Jim Head took on growing roles instructing Roosa, as well as the CM pilots on the final three flights, in the objectives of the photographic experi­ments and the cameras’ operation. After the film was returned, they also helped interpret the data obtained. Apollo 14 marked the end of the H missions, one short of the four originally planned.

The J Missions: We Almost Achieve Our Early Dreams

Apollo 15 was the first of the J missions. Years of struggle and cajoling, as well as long hours spent meeting with contractors, principal investigators, scientific committees, and NASA colleagues, had finally borne fruit. All the allowances for payload, extravehicular activity time, distance traversed, and sample return would suddenly double or triple. Although we had greatly increased our ability to explore, however, there would be no dual launches, no two-week exploration timelines to construct, and no seven-thousand-pound science and logistics payloads that would have given us the experience to plan for lunar bases. Our attempts to convince Congress and the Nixon administration to extend lunar exploration had failed. Instead of the five more missions we had been planning just six months earlier, only the three J missions remained. After Apollo 17 returned, Project Apollo would close its doors. We chose to put this sad ending out of our minds and concentrate on ensuring the success of the last missions.

Two days before the launch of Apollo 15 several of my colleagues and I flew to Orlando and then drove to Cocoa Beach, Florida, to prepare for the prelaunch press briefing. George Esenwein and Floyd Roberson would describe the new command and service module experiments; Ben Milwitsky and Richard Diller would do the same for the lunar roving vehicle; and I would cover the surface science. The routine at these briefings was that we would make short prepared statements, illustrated with vugraphs, and then take questions. Gene Simmons joined us, having recently transferred from MIT to the Manned Spacecraft Center, and some of the PIs—those with experiments flying for the first time— were on hand to discuss them. Gene, with his new title of chief scientist and an assignment to once again try to improve relations with the scientific commu­nity, had written a guidebook, ‘‘On the Moon with Apollo 15.” It was sought after by the media as a quick reference covering the science aspects of the mission and providing some easy quotations, a service the hardworking mem­bers of the press always appreciated. Gene compiled similar guidebooks for the last two missions, Apollo 16 and Apollo 17.

NASA’s Public Affairs Office usually released a mission press kit about ten days before the launch to give the media a chance to get familiar with the mission. It included details on all aspects, including the mission’s scientific activities and experiments. Before these last briefings at Cocoa Beach, media briefings for each launch were conducted at intervals at places such as MSC and Kennedy Space Center, and a major briefing was always held at headquarters about a month before launch for the large Washington media contingent. But the various briefings at Cocoa Beach (other parts of the flight besides the science were covered) the day before the launch were always the best attended, resulting in a lot of print and sound-bite coverage. Some seventy-five members of the media attended our briefing, held in a large second-floor conference room at the Friendship Inn the morning of July 25. Jack Hanley and Don Senich came along to answer any questions on two new pieces of equipment, the lunar drill and the soil mechanics penetrometer.

By this time our office staff had grown, mostly with members detailed from other agencies or NASA centers. NASA budgets had been going down for the past four years, and with those reductions came a semifreeze on hiring NASA civil servants. We always argued, to no avail, that some aspects of NASA busi­ness were still growing—Apollo science as an example—and needed more bodies. To spread the added workload in our office, we obtained detailees from the Army Corps of Engineers, the United States Geological Survey, and the Jet Propulsion Laboratory. In addition to Jack Hanley from USGS and Don Senich from the Corps of Engineers, USGS lent us Gerald ‘‘Jerry’’ Goldberg, and JPL sent Ewald Herr and later Peter Mason and Ronald Toms for assignments that lasted one year or more. Hanley, Goldberg, and Senich stayed with our office for over three years and were invaluable additions, cheerfully (usually) taking on the ‘‘dog work’’ that every government bureaucracy generates as well as the more interesting oversight for science payload development.

Besides the excitement of the upcoming launch, which brought media rep­resentatives from all over the world, in the days before the launch Cocoa Beach and the surrounding area were the site of many parties, a tradition that went back to the first rocket launches in the late fifties. These parties grew with each new manned launch. By custom, the night before the launch the big companies, with a major stake in the mission, would hold open houses that included food and drink. For Apollo launches, North American, Douglas Aircraft, Grumman, and Boeing, as well as smaller companies such as Bendix (many of these com­panies have since merged and lost their identities), would all hold their own affairs to tout their participation, with some competition to throw the best party. Similar “splashdown” parties were held in Houston, near MSC, after the end of each mission, and these would be even wilder, if that was possible, than the prelaunch parties at the Cape. These MSC parties were by invitation only, and invitations were always in great demand.

Up and down the beach a half dozen or more parties would go on into the wee hours. The morning after would be spent describing some of the more outrageous events. Nothing attracts the press more than a free party and, I might add, the many VIPs and sightseers who were in attendance. We civil servants tried to be discreet, but we would also drop in on a few of the parties even though some of us had official duties the next morning. With the Apollo 15 launch scheduled for 8:34 a. m., it meant waking up early to beat the traffic and get to my VOA broadcast site. I had been promoted since Apollo 11 and was now a full partner in the broadcasts.

The TV networks were getting more and more elaborate with their coverage of each succeeding mission. For Apollo 15, with its promise of real-time TV pictures during the astronauts’ LRV traverses, the major networks had assem­bled simulated lunar terrain to illustrate how the astronauts were going about their exploration. NBC had a small working model of the LRV, and CBS and ABC had borrowed full-scale working models. For TV and news media pools not fortunate enough to have their own models, we supplied static mock-ups of the LRV and Apollo Lunar Surface Experiments Package at MSC, where their reporters could be shown standing in front of the models.1 Each of the major networks also had a captive astronaut in the studio to explain the intimate details of what was going on during the mission.

In terms of science training, the crew of Apollo 15 was the best prepared yet. Based on my observations, Dave Scott, James Irwin, and Alfred Worden showed the greatest interest of any of the crews to date in understanding the science objectives for their landing site—the Apennine Mountains and nearby Hadley Rille—and the new suite of experiments housed in the lunar module and CSM.

Scott, as mission commander, set an example for his two crewmates through his hard work and contagious enthusiasm. Hadley Rille, a long, sinuous valley, was one of the most intriguing features on the Moon’s surface. It was surpassed in interest only by the first landing site, when any information returned was bound to be extraordinary. Several theories had been proposed to explain the rill’s origin: that it had been formed by water discharged from the Moon’s interior (the least favorite theory among most lunar scholars); that it was a lava channel or collapsed lava tube or had been gouged from the surface by some volcanic event; or that it was the remnant of faulting or stretching of the lunar crust.

The landing site was on the eastern rim of the Imbrium basin, so we antici­pated that the returned samples would include material from the Apennine Mountains, probably formed by uplift and ejecta from deep in the Moon’s interior. Other samples should include Imbrium basin fill consisting of some type of lava. Samples from the edge of Hadley Rille might resolve the question of its origin. We might even collect samples of the ejecta from Mare Serenitatis, just a short distance to the east. If we could identify their source, age dating these samples would go a long way toward explaining key events in the Moon’s early history that shaped its final form.

Adding to our excitement about this mission was the greatly increased radius of operation for the astronauts and the many new experiments that would be performed. The science payload delivered to the lunar surface would be almost 1,200 pounds, compared with the Apollo 11 payload of less than 200 pounds and more recently the 470 pounds carried on Apollo 14. If the mission went as planned, Apollo 15 would come closest to our earlier dreams for the first post – Apollo missions. Apollo 15 marked another milestone: Jack Schmitt was named to the backup crew. Based on previous crew rotations, this would have put him in position to be named to the prime crew for Apollo 18, now canceled, but at least he had moved up in the pecking order. This would be Gordon Swann’s last mission as PI for the field geology experiment. He and his many Flagstaff colleagues and coinvestigators, which included Bill Muehlberger, the designated PI for the last two missions, had been building to this climax after many years of hard work. Next best to being on the Moon themselves, the J missions would validate their efforts, and they worked tirelessly to prepare the crew.

All the preparation paid off. The crew performed flawlessly. The Science Support Room (SSR) geology team, led by Swann, listened, recorded, debated, and attempted to interpret in real time everything that was happening on each of the three EVAs during the 67 hours the crew was on the surface. Total time of the three EVAs was 18.5 hours, a new record. As we had practiced at Flagstaff, Scott also performed the first and only stand-up EVA when, shortly after land­ing, he opened Falcons overhead hatch and stood on top of the ascent stage engine cover to get a bird’s-eye view of the landing site. While enjoying the scene around him, he took some panoramic pictures and planned the upcom­ing traverses. With only minor upgrades of LM systems for the J missions, we were able to more than triple the time the Apollo 11 crew had spent on the lunar surface. The Apollo 15 EVA time was almost as long as the total time the Apollo 11 astronauts spent on the Moon. These numbers confirmed in my mind that the one – to two-week visits we envisioned for the post-Apollo dual-launch missions could have been achieved by making the modifications to the Apollo systems we had studied.

Another change for Apollo 15 was the inclusion of scientist-astronauts Jack Schmitt, Joe Allen, Bob Parker, and Karl Henize as capsule communicators; Allen, who had served as mission scientist, usually manned the console during the EVAs. On all the previous missions only Schmitt, for Apollo 11, had been given this high profile task. Although it is difficult to point to any specific advantages of having them at the consoles, interaction between the ground and the crew of Apollo 15 was lively, and certainly we in the SSR felt more comfort­able knowing that Allen could immediately interact with the crew if necessary. We passed suggestions and questions to the CapComs, and in contrast to Apollo 12, many were passed on.

The scientific harvest from Apollo 15 was spectacular, derived both from the lunar surface and from lunar orbit. An ALSEP was deployed at the north­ernmost point reached by any Apollo mission. This ensured an ideal position­ing of the ALSEPs for triangulating readings for experiments like the passive seismometer and LRRR that needed site separation and the surface magne­tometer that was attempting to discover if the Moon’s magnetic field might vary from site to site. The three LRV traverses covered almost seventeen miles, during which the astronauts studied twelve locations in addition to the imme­diate landing site. They collected almost 170 pounds of samples and took more than 1,100 photographs. Besides the many photographs taken with the Hassel – blad cameras, we also had TV coverage of eight stations and some footage taken while the astronauts were under way on the LRV. Eight major experiments were conducted in lunar orbit, and many types of photographs were taken from both the CM and the scientific instrumentation module (SIM) bay in the service module, adding to the wealth of new information that included coverage of the Moon’s farside.

Apollo 15 recorded one other first, and last. After Scott and Irwin’s ren­dezvous with Worden back in lunar orbit, Lee Silver was called out of the SSR to go to the Mission Operations Control Room. Lee had led many of the field training exercises for the Apollo 15 crew and had established a close rapport with them. Scott wanted to talk directly to Lee to thank him for his long hours and dedication to their education. Thus, for the first and last time during an Apollo mission, a member of one of the science teams talked directly to the astronauts while a mission was in progress without going through an astronaut CapCom. Lee passed on our congratulations and told them how excited we were about what they had accomplished, and he reported the results we were already seeing as we reduced the traverse data. I hope that when we return to the Moon such exchanges between scientists on Earth and those working on the lunar surface will be the norm, for it will surely add to the value and efficiency of future lunar exploration.

All the experiments and equipment carried on Apollo 15 performed up to expectations except for the drill. To quote from the crew’s observations, ‘‘The deep core could not be extracted from the uncooperative soil by normal methods; the two of us, working at the limit of our combined strength, were ultimately required to remove it.’’ The exterior flutes contributed to this condi­tion because the drill stem was pulled into the ground still deeper when the motor was activated.2 This activation was supposed to clear the flutes for easy extraction.

As we were to discover, there were two complications involving the drill. The first and most serious occurred as the astronauts were drilling the bore holes for the heat flow experiment, and they encountered the second while trying to extract the core sample. Mark Langseth, the heat flow PI, had designed his experiment around placing the sensors in a cased drill hole some ten feet deep. When Scott attempted to drill the first hole he could not go much deeper than about five feet, well short of his target. The drill stem refused to go any farther no matter how hard he tried. Frustrated and thinking he might have hit a large rock, he stopped working on the first hole and tried to drill the second. Again, he could not get penetration much below the first length of drill stem. Time was fleeing, so he was instructed to stop drilling, place the sensors in the holes he had, and finish the remaining first EVA tasks. Langseth ended up with his sensors much closer to the lunar surface than he wanted, and he feared he would not get the high quality data he was hoping for. More later on the tribulations of the heat flow experiment.

In drilling the core sample the astronauts encountered a different problem. The drill penetrated to the full depth quite easily—too easily, it turned out. Fortunately, through the crew’s ‘‘combined strength’’ they salvaged the core. Thus the time had not been wasted. Wasting time during a mission was not attractive to anyone, especially when the medical team monitoring the astro­nauts could see they were in danger of exceeding their physical limits while trying to remove the core. As the astronauts struggled, recommendations were made in Mission Control to abandon the attempt, but Scott and Irwin persisted and saved the day.

Immediately after this well-publicized glitch, while the crew was still on the Moon, Rocco Petrone caught me in the Mission Control Center and issued one of his famous edicts—we were to solve the drill problem before the Apollo 16 flight readiness review! I huddled with Jack Hanley to discuss our course of action. Following our usual method of addressing such issues, we appointed a ‘‘tiger team,’’ consisting of Dave Carrier and several other MSC engineers plus Jack Hanley and Don Senich from my office. They were dispatched to Denver to meet with Martin Marietta, find out what went wrong, and make the necessary modifications. The solution had to be found quickly, since the drill had to meet the Apollo 16 equipment stowage window for its preflight checks. We also had to be prepared to modify the astronauts’ training and simulation schedules if any changes required new instructions or training.

After discussing the crew’s observations and reviewing how the drill had performed, we knew we had two separate problems, one in drilling the bore holes for the heat flow experiment and the second in extracting the deep core. The tubular drill sections used for the heat flow holes were of a different design than the core stems. Since the core stems had drilled to almost eight feet without any trouble, this design difference had to hold the clue. Langseth thought he knew what had gone wrong. The flutes on the outside of the heat flow drill sections did not extend the full length of each section; they stopped short of the ends, leaving an open space on the shafts.

With new information on the characteristics of the lunar soil derived from the soil mechanics experiment, the fidelity of the simulated lunar soils used during testing was improved, and we conducted many tests, some in a vacuum chamber. Langseth was right. The short interruption of the flutes at the joints had prevented the cuttings from traveling up the tubular sections; they jammed at the joint after the next section was added and the joint was drilled a short distance below the surface. The design had worked satisfactorily during our terrestrial trials, but the soil simulant used during the original tests had not been as compact and dense as real lunar soil.

The original design of the heat flow tubing had called for titanium inserts at the ends of the fiberglass sections to strengthen the joints so they could be screwed together like the core sections. This would also ensure good meshing of the flutes, but Langseth had been concerned that so much metal in the tubing might disturb the sensitive readings he was hoping to obtain. We had removed the titanium joints, and the tubing carried on Apollo 15 required that each section be pushed into the next. The flute alignment was determined by how carefully the astronauts joined the sections, but there was always a small gap be­tween the flutes. Langseth agreed that we would have to go back to the original design for Apollo 16. This was done, and the drilling tests were successful.

We believed we had solved the heat flow drilling problem, but we needed more tests to understand the core extraction problem. During the crew debrief­ing it emerged that because Scott had had trouble drilling the heat flow bore holes, he had perhaps put too much pressure on the drill while coring. This forced the core stem into the soil before the flutes could completely clear the cuttings, thus jamming the core stem. Tests at Martin Marietta showed that if the drill penetrated more slowly, even in a more moonlike soil than we had used in previous simulations, there should be no difficulty extracting the core on the next missions. Just in case it did jam, we designed a new device to jack the drill stem out of the hole if it would not come free using the normal procedure of rotating the drill core in place. Score two hits for the tiger team and the Martin Marietta and Black and Decker engineers. On Apollo 16 and Apollo 17 the drill worked well for both applications.

Although the missions, starting with Apollo 15, were being launched on a more relaxed schedule, approximately one every eight months, there was a heavy training burden on everyone at MSC, KSC, and the Field Geology Team. At headquarters we also felt the pinch as we tried to stay abreast of the progress, or lack thereof, so we could keep Petrone and other senior management up to date. Rocco hated surprises, and this attitude carried over to all his staff and was often reflected during his weekly status reviews. These reviews, held at our offices at L’Enfant Plaza in a large, windowless room lined on both sides with multiple sliding status boards, would consume half a day or longer depending on the number of outstanding issues. The status boards were updated daily through a contract with the Boeing Company and were used extensively during the reviews. Each Apollo office would make a presentation so that Rocco could get a snapshot of the program covering everything from spacecraft and payload status to funding, manpower, and eventually, final plans for close-out. After the near disaster of Apollo 13 and with the last missions firmly scheduled, the atmosphere was getting tenser; we had to make sure that nothing was left to chance and that no dumb mistake would jeopardize a crew.

In November 1971, four months after the return of Apollo 15, Lee Scherer was transferred to a more prestigious management position, director of the NASA Dryden Flight Center at Edwards Air Force Base in California. Don Wise, Lee’s deputy, tiring of the Washington scene, had already gone back to academia. O. B. O’Bryant was named to replace Lee for the final two missions, and I was named program manager for Apollo surface experiments, including the ALSEP, since Ed Davin and Dick Green had left to take new jobs at the National Science Foundation.

The crew of Apollo 16—John Young, Charley Duke, and Kenneth Mat­tingly—had been named long before Apollo 15 was launched, and their training and simulations overlapped those of the Apollo 15 crew. For the Field Geology Team there was the added consideration of changing the PI from Gordon Swann to Bill Muehlberger, with Bill adding some new coinvestigators from USGS and Bellcomm. This transition went off without a hitch, since both Gordon and Bill had worked together for a long time and there were no profes­sional jealousies involved in the switch. With the overlap in training the Apollo 15 and Apollo 16 crews, it was difficult even to notice a change, and almost all the faces remained the same. With each mission, the training was also becom­ing more complicated, reflecting the added complexities of the J missions with their new experiments, both surface and orbital, and longer surface EVAs.

The Apollo 16 crew had a much different character than the crew of Apollo 15, reflecting the personality of John Young. He had already flown three space missions, including the highly successful Apollo 10, and had more time in space than any other astronaut except Jim Lovell. Young was more relaxed than the hard-driving Scott. He was quick with a quip or story to break the intensity of a training or simulation session. In spite of the more relaxed atmosphere, the crew was required to spend hundreds of hours learning the scientific nuances of their chosen landing site, Descartes, a highlands crater near the Moon’s center several hundred miles southwest of the Apollo 11 landing site. In addition to learning what they should expect and look for at Descartes, they had to train to deploy the ALSEP and all the other experiments, including the redesigned drill and one new experiment, the far UV camera-spectrograph, our first chance to use the Moon for astronomical studies.

Bill Muehlberger, perhaps tempered by dealing with undergraduates during his tenure at the University of Texas, Austin, plus his long association with some of the eccentric personalities at USGS, meshed well with the crew. His team of coinvestigators and field geology instructors took on the task of instructing the crew by taking them to several training sites in the United States, Mexico, and Canada. The sites were chosen to expose them to field conditions representative of the latest geologic interpretations of what they might encounter at Descartes. Since this would be the first landing in the Moon’s highlands, we expected to pin down their composition, an important determination in understanding the Moon’s history. The crew would also sample the material that filled the large crater in which they would land; perhaps it was of a different composition than the maria sampled on the earlier missions. Some photogeologists studying the area around the landing site believed they were observing volcanic features, low hills that might be composed of lava or cinders, much like the formations just east of Flagstaff. Other interpretations were possible, but these hills looked unusual and, most important, were in the highlands. Young and Duke were conscientious students and quick learners. Those of us who tagged along to observe the training sessions were impressed at how well they were absorbing the huge amount of information thrown at them. The sessions included a trip to Sudbury, Canada (which I didn’t attend), considered an excellent example of some of the geological situations they might encounter on the Moon.

Ken Mattingly, the CM pilot, who had been scratched from the Apollo 13 crew because of fears he might have been exposed to German measles, for which he had no immunity, was the most studious of the three crewmen. He realized how fortunate he was to have a second chance, and he was determined to get the most out of the experiments he would operate from the CSM. Like Al Worden, he would be photographing and making measurements over a wide swath of the lunar surface. He spent many hours with Farouk El Baz, Goddard Space Flight Center PIs Isadore ‘‘Izzy’’ Adler and Jack Trombka, and other PIs, learning as much as he could about their experiments and what they hoped to achieve. The results of the Apollo 15 orbital science were now available. Draw­ing on Worden’s experience of operating the experiments for the first time, Mattingly was in a better position to manage them efficiently.

Apollo 16 was launched from KSC just before 1:00 p. m. on April 16, 1972, a more civilized hour for those of us covering the launch. There were no com­plications with the flight until after lunar orbit was achieved, but a major problem surfaced just after the LM and CSM separated. When Mattingly went through his checklist before firing the service module engine to circularize his orbit—his first order of business after separation—one of the gimbal motors that controlled the SM engine nozzle did not respond properly. If he could not get the gimbal motor to work, he would not be permitted to start the engine. Mission rules dictated that the landing would have to be aborted and the LM would rendezvous with the CSM using the LM ascent or descent engine or both. Once joined, LM propulsion would be used to get them out of lunar orbit and on the way back to Earth.

While Mission Control tried to find a solution, the LM and CSM were directed to orbit the Moon near each other but not to join up. The crew, and all of us sitting on the edge of our seats in the SSR, kept hoping the mission would not have to be aborted, but with every orbit that passed without instructions on how to proceed, it was becoming less and less likely that the landing would happen. If the landing was permitted, time lost would have to be deducted from the lunar surface staytime. Eventually, if too much time elapsed, the landing would have to be called off because the sun angle would impair visibility on the surface so Young would be unable to avoid small obstacles at the landing site.

After reviewing the data sent back to Earth from the CSM and consulting with the North American engineers, Mission Control decided it was safe to proceed. Mattingly successfully fired the engine to put him into the desired circular orbit, and Young and Duke completed their landing. The landing delay (approximately 5.75 hours) did reduce the time spent on the lunar surface. To make up for this lost time it was proposed to cancel the third EVA. After much pleading from Muehlberger’s team, emphasizing the importance of the sam­pling sites selected for the third EVA, it proceeded as scheduled but was reduced by about two hours. To keep on the overall flight schedule, the time would be

made up by lifting off from the Moon sooner after the astronauts returned to the LM at the end of the third EVA than originally planned.

With their problems behind them (there were a few others), Young and Duke went about exploring their landing site and deploying all their experi­ments. The ALSEP was set up during the first EVA, and this time the drill worked as designed. But the heat flow experiment met with calamity. After the first hole was drilled and the probe was lowered into the tubular casing, Young, working on another part of the ALSEP deployment, tripped over the cable connecting the probe with the central station and pulled it loose. ALSEP cables consisted of copper wires embedded in a thin plastic covering several inches wide, designed to lie flat on the lunar surface after they were unreeled from their containers. But after being coiled for several weeks in stowage, they tended to develop slight kinks. Whether that caused Young to catch the cable with his boot or whether he just misstepped, the cable came loose. After examining the end of the cable that had been torn off and describing it to Mission Control, it was decided it could not be reconnected.

We never anticipated a failure of this type, so no tools were carried for making a repair. Just to be sure we weren’t overlooking a possible fix, and with a distraught PI begging us to find a solution, we put together another tiger team. But we could not come up with a guaranteed way to reattach the cable. With the cable broken the experiment could not operate, so we canceled the drilling of the second hole. This was a major setback for Langseth; the first deployment of his experiment had placed the probes too shallowly, compromising the read­ings, and this deployment was a complete failure. His experiment, considered one of the most important we would place on the Moon, would have one last chance on Apollo 17. But with the loss of this data point and the compromised data from Apollo 15, even if the Apollo 17 deployment was successful our overall understanding of the heat flow from the Moon’s interior would be open to question, since the measurements at the Apollo 17 site might not be typical for the Moon as a whole.

The deep core was drilled successfully on the first EVA, and eight feet of core were recovered. The active seismic experiment, a duplicate of the one deployed during Apollo 14, functioned much better this time. We had made some minor changes in the thumper firing mechanism, and all twenty-one charges fired. The second part of the experiment, the mortar package, was placed in a better position relative to the ALSEP, and we were able to fire three of the four mortars one month after the astronauts returned home. After firing the third mortar, the pitch-angle sensor showed that the mortar box might have tilted, so we decided to hold off firing the fourth mortar until later.3

The far UV camera-spectrograph, the first telescope to be used on the Moon, was placed in the shadow of the LM and moved several times on succeeding EVAs to keep it in the shadow. It was pointed at different sectors of the sky, three times during the first EVA, four times during the second, and three times during the third, then the film was unloaded and returned to Earth. After setting up the ALSEP and other experiments during the first EVA, the astro­nauts returned to the LM and unloaded the LRV. This left just a short time for the first traverse, and they went to study and sample some craters less than a mile to the west.

On the second EVA the astronauts traveled south about two miles with a major objective of sampling the debris thrown out by a ‘‘recent’’ impact crater (named South Ray) about half a mile in diameter that had scattered ejecta a great distance in all directions. It was expected that this ejecta would provide us with good samples of a geological formation (named Cayley by USGS) that forms extensive highland plains thought by some to be volcanic in origin. If it proved to be volcanic, its composition and age would be important pieces of information for understanding the development of the lunar highlands that make up approximately four-fifths of the Moon’s surface. Equal in importance to resolving the composition of the Cayley formation was obtaining samples of the mountain-making material in the vicinity of the Descartes Mountains. Although the mountains were some distance from the landing site, we believed we stood a good chance of recovering rocks deposited in the plains from impacts that occurred in the highlands.

Tony England, selected in the second scientist-astronaut class, was the mis­sion scientist and CapCom during the EVA periods, and he did a good job of communicating with the crew and relaying questions and suggestions from the SSR. In his role as mission scientist, he had accompanied the crew on many of their training trips and participated in the simulations leading up to the mis­sion, as Joe Allen had for Apollo 15. He was intimately familiar with all the equipment and experiments and was able to quickly give advice when needed.

SSR operations had improved with each flight. Beginning with Apollo 15, we could supply more and more backup information. We kept careful track of the EVAs as they progressed; the planned traverses with each station were identified on a three-dimensional model constructed from the Lunar Orbiter photo­graphs, and we had a list of planned activities for each stop. We also kept a traverse profile showing in graphic form the status of the EVA in terms of time and life-support expendables. Using this profile, we were prepared to suggest modifications to the EVA plans if something unexpected happened or if the astronauts spent more time than planned at a given station. Time was always our enemy, and we knew the crews felt the same. How could they make the most of each minute yet not miss some important discovery? With the pictures coming back from the TV camera carried on the LRV, we were able to keep up with the crews’ efforts and think ahead with them as to what should be the next priority. However, the crews always seemed to make the right decisions without many inputs from the ‘‘back room,’’ a tribute to their training and dedication. There was seldom any second-guessing from those of us privileged to feel so close to the action even though we were 238,000 miles away from actually swinging a hammer or snapping a camera shutter. They were our surrogates in this inhospitable and strange land; we could only sit back and admire the job they were doing. Our job was to try to assimilate the information they were returning so as to arrive at the ‘‘big picture’’ of the Moon that their observations were starting to give us. We were careful not to interrupt the crews or burden them with unneeded questions.

The third EVA that the Field Geology Team had pleaded to retain became a quick trip to sample a northern crater (North Ray), about three miles away. Deposits around this crater were believed by many, but not by all, to be the best opportunity to collect volcanic samples. Time permitted only a few stops, but the traverse was made without incident and harvested many samples. The total weight of samples collected during the three EVAs was 211 pounds, a new record; the total distance traveled, sixteen miles, was slightly less than that recorded for Apollo 15.

Ken Mattingly, after the rocky start caused by the spurious gimbal motor readout, had successfully carried out his part of the science tasks. His sensors and cameras covered much of the Moon’s surface between five degrees north and five degrees south of the lunar equator. Ken also spent as much time as possible making careful visual observations, which were valuable during the crew debriefings and in analyzing the data captured by the sensors.

Several months later, analysis of the Apollo 16 samples showed that they, like most of the samples collected from the previous four missions, were breccias, the products of one or more cataclysmic events that showered the lunar surface with the debris from multiple impacts. None were volcanic in origin, however, and the Lunar Sample Preliminary Examination Team stated that ‘‘no evidence for lava flows or pyroclastic rocks was observed.’’ This was an example of the pitfalls of trying to make photogeologic interpretations with no fieldwork to base them on. But we were learning with each mission how to improve our interpretations, and now we had an additional tool, the CSM data collected from lunar orbit that provided estimates of the composition of the lunar sur­face over wide areas.

With the Apollo 12 ALSEP’s third anniversary of uninterrupted operation approaching, I wrote a memo for Rocco Petrone’s signature that was distributed to NASA management, including the new NASA administrator, James Fletcher, to report how well the ALSEP had performed.4 We reminded those on the distribution that the original design goal was one year and that four experi­ments, the passive seismometer, Suprathermal Ion Detector Experiment, Solar Wind Spectrometer, and dust detector, were still operating normally and re­turning useful data. The Lunar Surface Magnetometer had operated success­fully for two years, and only one experiment, the Cold Cathode Gauge, had failed immediately after activation.

The radioisotope thermoelectric generator was still putting out sixty-nine watts of power, four watts above predicted values for initial power output. The ALSEP central station had responded to over fifteen thousand commands dur­ing the three years and showed no sign of deterioration in spite of having experienced thirty-seven lunations that created temperature swings each time of over 500°F (— 260°F to 270°F; by this time we had a more accurate measure­ment of surface temperatures). The Apollo 12 ALSEP would continue to operate for almost five years longer.

It was 9:00 p. m. on December 6, 1972. Apollo 17, the last of the Apollo lunar flights, was on launch pad 39A at KSC with my good friend Jack Schmitt, Commander Gene Cernan, and CM pilot Ronald Evans strapped on board the command module, named America. As we had done for the last two flights, Rhett Turner and I were prepared to broadcast the launch for the Voice of America Worldwide Service from the press site about three miles west of the pad. To our right, along the raised, curved berm, were the large air-conditioned broadcast booths of CBS, ABC, and NBC, along with booths of other com­panies. If we used our binoculars we could see Walter Cronkite, Jules Bergman, and the other TV commentators looking out their picture windows at the brightly illuminated Saturn Vi Voice of America was a bare-bones operation. We sat in the open swatting mosquitoes, last in line on the berm, on two folding chairs at a card table. Behind us, in what was once a two-wheeled camper trailer, was the engineer with all the electronic equipment. Later, after the lunar landing, along with my regular duties in the science support room at MSC, I was scheduled to make several broadcasts for the Spanish-language VOA using my rusty Colombian Spanish to explain what was happening—my last assign­ment for VOA.

Apollo 17 was scheduled for launch at 9:53 p. m., the first night launch for Apollo. So in addition to our excitement about this last launch and all it meant, we were looking forward to seeing and ‘‘feeling’’ the giant Saturn rocket roar off the pad. We could only guess at the visual impact, the spectacle of the world’s largest firecracker lighting up the night sky. As we sat watching the brilliantly lit launch tower and rocket, the clouds that had partially obscured the sky were slowly dissipating; it had been predicted that if the sky was clear viewers as far as five hundred miles away might see the rocket as it streaked away to the east. Rhett had done his usual impeccable homework for the launch, the best of any reporter I knew, and he carried the audience along as the countdown pro­ceeded, bringing me in as needed to provide some special insight. As we went through our rather informal script, Rhett would signal the engineer to play a previously recorded interview with Jack, Gene, or Ron or some other pertinent clip that would give interesting background about the mission. Illuminated in front of us, between our position and the pad, was the large digital countdown clock counting down the seconds. From time to time Chuck Hollingshead, the voice of Kennedy launch control, would interrupt our coverage with comments broadcast over the public address system.

Everything was proceeding normally until T minus thirty seconds, when without any warning a hold was announced. No immediate reason was given, and we were left, with all the rest of the commentators, to speculate on what was wrong. For the next twenty minutes we tried to make educated guesses, hoping that it was something minor and the countdown would soon resume. Finally Hollingshead came on the PA speaker and explained what caused the hold. The third-stage fuel tanks had not pressurized on schedule, and though a manual pressurization was attempted it was too late in the countdown and the auto­matic sequencer shut down the launch. It didn’t sound serious, but it wasn’t clear when the count would start again; we assumed it might take about an hour before the countdown would resume at T minus twenty-two minutes, the newly announced recycled starting point.

We underestimated. For the next two hours Rhett and I filled the airways with impromptu discussions of Apollo 17 science and whatever other subjects we could think of. From time to time VOA would break away to provide news of the world and return to us, still sitting under the stars waiting for an announce­ment. Listening to those VOA tapes twenty-five years later is a real trip down memory lane. They record the late-breaking news and include the story that former president Harry Truman, age eighty-eight, was in critical condition in a Missouri hospital.

Finally the count resumed at T minus twenty-two minutes; it was held again at a planned point at T minus eight minutes to check that the pressurization trouble had been resolved before the countdown continued. At 12:33 a. m. on December 7, 1972, Apollo 17 was launched; a historic date, one that will always be remembered along with another December 7 thirty-one years earlier. The liftoff was every bit as spectacular as we had hoped, lighting the night sky for miles around and pounding our bodies with the powerful low frequency rever­berations that only a Saturn У launch produced. If you have witnessed a shuttle launch, multiply the effect by two. The crew of Apollo 17 was on its way, and to top it off, we had survived two hours of unscheduled airtime! We packed our notes and left for Houston.

Apollo 17, after its prelaunch difficulties, was the most trouble free of any of the missions. All the pieces were falling neatly into place. We were definitely learning, but now we had no further chance to put this hard-won education to use. The landing site, Taurus-Littrow, almost as far north as the Apollo 15 landing, was on the edge of Mare Serenitatis, to an Earth observer the right edge of the man in the Moon’s right eye. The landing would be the most difficult maneuver of any flight yet, requiring Cernan to come in over the Taurus Moun­tains, 6,500 feet high, descend steeply into a narrow valley, and land between the bases of two mountains.

Taurus-Littrow was selected for the final Apollo landing for several reasons. During the Apollo 15 mission Al Worden had observed that this area was covered by a mantle that looked darker than other parts of the lunar surface. His observations seemed to be confirmed by Lunar Orbiter photography and photographs taken from orbit during other Apollo missions. The promise of finding “recent” volcanism raised its head again. Would this site provide sam­ples that would confirm an epoch of late lunar volcanic activity? Samples from the Taurus Mountains were also of great interest. Would they be similar to or different from the highlands samples collected on Apollo 16?

As a far northern and eastern landing site, it had value for several of the ALSEP experiments, in particular for the Lunar Atmospheric Composition (LACE) experiment (the passive seismometer was not carried on this mission), which would provide better data if separated in distance from the other ALSEPs, which were still sending measurements. LACE, a miniature mass spectrometer, was a more sophisticated version of the Cold Cathode Gauge experiment de­ployed on missions 12, 14, and 15 to detect the tenuous lunar atmosphere. We also had two new surface experiments on the mission, Surface Electrical Proper­ties (SEP) and the traverse gravimeter, which, along with the portable magne­tometer that flew on Apollo 16, would be operated by the astronauts during the LRV traverses. These last three experiments were expected to provide important information on the subsurface structure of the valley at the base of the Taurus Mountains.

While missions were under way, my job at MSC included manning a console in the Science Support Room and taking part in the discussions that would fill the exciting hours while the astronauts were on the lunar surface. Occasionally I would spell the headquarters duty officer in the Mission Operations Control Room, the latter largely a ceremonial duty if the mission was proceeding ac­cording to plan. In addition, I would participate in briefings with VOA and other news organizations. Although the word ‘‘spin’’ had yet to be applied to government briefings, that was part of our approach. If something in the mission timeline didn’t go according to the material passed out to the media before the mission, we would be interrogated at each of the daily updates held in the MSC auditorium. No question was off limits, and some would be off the wall, reflecting the media’s understanding, or misunderstanding, of what was going on.

We all had our preferred media person to talk to off line, someone we knew from experience would tell the story reasonably straight and get the facts right. My favorite was Donald Kirkman of Scripps-Howard. I tried to avoid Thomas O’Toole of the Washington Post, with some success, for he seemed to be always looking for the negative side of events and could usually be counted on, at some point during a mission, to misinterpret an important story. We would feed trusted reporters tidbits of insider information so that their stories would be more informative or have a little more punch than their competitors’.

Just before the crew achieved lunar orbit, their discarded SIVB stage hit the Moon about 525 miles west of the Apollo 16 ALSEP, and the impact was re­corded by all four passive seismometers that were still operating from the earlier missions. This time there were no problems in lunar orbit after separa­tion of the LM and CSM, and Cernan accomplished the landing after taking over control from the autopilot and set the lunar module, Challenger, down in the rock-strewn valley between the North and South Massifs.

Once on the surface, Cernan and Schmitt, the last men to set foot on the Moon for what has turned out to be three decades and counting, went energet­ically about their business. The crew had trained hard, and we could sense from Cernan’s descriptions that he was not about to be outshone by his geologist teammate when it came to conducting their surface studies. They described the sight that confronted them as “spectacular.” I have probably overused that word in this story as much as the astronauts did during the missions, but it is the best adjective I know to describe the views we could see from the TV images and later from the many excellent photos they returned. To the north, less than seven miles away, the mountains rose almost perpendicular from their base toward the black sky. To the southwest, again less than seven miles away, lay the South Massif, equally imposing if not quite as steep as the mountains to the north. TV pictures captured the landscape clearly, and in the SSR we could only wonder at our audacity in asking the crew to land in such close quarters.

Apollo 17 reconfirmed the targeting ability of the MSC engineers. They brought Cernan and Schmitt to the precise point where Cernan was scheduled to take control, and he then successfully demonstrated his landing skills. With this experience it seems certain that if missions had been scheduled after Apollo 17 we could have persuaded management to agree to landings at important sites such as the central peaks or rims of Copernicus and Tycho. Future lunar explorers, undoubtedly piloting spacecraft with greater capabilities, will find safe landing sites almost anywhere on the Moon, including the farside!

Apollo 17’s first EVA began with the removal of the LRV from its stowage bay on the descent stage and the erection of the TV high gain antenna. Thereafter we had good TV coverage of the landing site and the astronauts deploying the ALSEP. They drilled three holes, two for the heat flow experiment and the third to recover a ten-foot core. ALSEP began transmitting data as soon as it was activated, and the star of the experiments, the Surface Gravimeter, provided strong signals. Little did we know at this point that there was a major problem, as described in chapter 7. After finishing these tasks, the astronauts still had time to take a short (about half a mile) ride to the south to collect samples near a small crater named Steno. During this traverse they also took Traverse Grav­imeter and SEP readings and left two explosive charges to be detonated later to provide signals for the Seismic Profiling experiment.

A few words about the explosive packages that were an integral part of the Seismic Profiling experiment. Commencing with the design of the active seis­mic experiment, carried first on Apollo 14, we went through an extensive review and certification of the explosives used with that experiment and the Seismic Profiling experiment. Some at NASA were not happy about carrying live explo­sives on the LM, so our test procedures were carefully monitored. We had to prove beyond any doubt that there could be no accidental firing of the charges. Petrone, especially, followed the certification process from beginning to end and witnessed some of the field tests.

Fortunately this experiment was not the only place explosives were used during the mission, starting with the separation of the launch escape tower from the CSM and progressing through the individual rocket stages, where explosive squibs were used to separate some of the stages during flight. We benefited from all the work that went into qualifying these explosives and designed our charges using aspects of these proven designs. The biggest fear, of course, was that an inadvertent firing command, short circuit, or other acci­dent might trigger the explosives, either while they were stowed on the LM or while the astronauts were setting up the experiments.

Because the astronauts would hand carry the explosives on the lunar surface, every firing circuit had either double or triple safety redundancy before the firing commands could activate the charges. For the Seismic Profiling experi­ment, the arming sequence was as follows: Each explosive package had three pull rings on top. Pulling ring one started the safe/arm timer. Pulling ring two, and rotating it ninety degrees, released the safe/arm slide to start the mechan­ical timer. Pulling ring three cleared the firing pin and placed a thermal battery timer on standby until a coded signal was received from the ALSEP central station, the preferred way to set off the charges. In case ALSEP commands weren’t received, the mechanical timers were preset for periods from 89.75 to 92.75 hours after activation, well after the astronauts left the lunar surface. Each charge package had an antenna that would receive the initiation signal from the central station to start the firing sequence, at which point there was a two – minute window in which to receive the coded firing signal.

Sound complicated? It was. There was a running joke that with all the safety features we would be lucky to get even one to fire. But all eight charges were fired successfully after the astronauts departed, and the experiment’s four geophones recorded the explosions, providing information about the upper mile and a half of the Moon’s subsurface. The LM ascent stage impact, about seven miles southwest of the landing site, with its higher energy input, allowed Bob Kovach, the PI, to improve his measurements and estimates of seismic velocities to a depth of three miles below the surface.

The second EVA, the longest of the three, went almost due west and then swung southwest to study and sample the South Massif. Cernan and Schmitt made several stops along the way, including sampling the rim of a fairly large crater called Camelot. On finishing their work at the crater, they drove up the low scarp that separated the valley from the massif and sampled the boulders at the base of the massif. They then returned to the LM by a different route that took them farther north on the valley floor and included seven more sampling stops. One of these stops, at Shorty Crater, was to sample the dark halo material surrounding the crater that could be seen on Lunar Orbiter photographs. Worden had reported that he could see a color difference from orbit. Once again we hoped the dark material would be ‘‘recent’’ volcanic deposits as was predicted (incorrectly) at the Apollo 16 landing site. What they found, to the great excitement of both astronauts, was an orange and red soil interspersed with darker and lighter soils. Schmitt thought they had found the elusive recent volcanic vent. Returning to the LM at the end of the EVA with their find, they placed three Seismic Profiling explosive packages at varying distances from the ALSEP central station and took a series of Traverse Gravimeter and SEP readings.

The third EVA traverse was made toward the east and then turned north to sample the North Massif and the intervening darker plains material, also inter­preted as volcanic mantling material. Traverse Gravimeter readings were made on this EVA, and the final three Seismic Profiling explosive charges were placed on the surface at intervals along the route. Because the SEP receiver overheated, no data were collected by this experiment during the third EVA.

By now, after analyzing the signals coming back for almost two days, we realized that the Surface Gravimeter was not responding correctly. We cut the third EVA short to allow Schmitt to go back and rebalance the gravimeter’s movable beam. This last attempt to improve the experiment’s response while the astronauts were still on the Moon also failed. Something was wrong, but we weren’t sure what it was and couldn’t find a solution. Closing out the final Apollo lunar surface EVA on a somewhat dismal note because of the gravimeter problem, Cernan removed the neutron probe from the core hole for analysis back on Earth and climbed back into the LM.

With the crew of Apollo 17 safely on the Moon, Dale Myers and Rocco Petrone released the final ‘‘Apollo Program Plan.’’5 It covered all the remaining Apollo 17 activities and those actions necessary to close out the Apollo program. Among other notes of finality, it stated: ‘‘All basic hardware procurement for the Apollo Program has been accomplished.’’ The nation would not purchase any more awe-inspiring Saturn Vs or superbly engineered CSMs and LMs. The schedule in the plan showed a transfer of responsibility for common Apollo – Skylab activities to the Skylab Program Office, the next approved program, by the middle of FY 1973. Beginning on the same date, the remaining Apollo lunar science activities, mostly monitoring the ALSEPs and publishing results, would be undertaken by the Office of Space Science and Applications. Although Apollo 17, and for that matter the entire Apollo program, had achieved all its objectives and more, this final Apollo plan ended an era with the sour taste of a great opportunity lost through lack of national leadership. It was an era that had begun with great expectations of conquering new worlds.

After almost seventy-five hours on the lunar surface, twenty-three of them spent outside the LM on the three EVAs (another new record), Cernan and Schmitt lifted off to rendezvous with Ron Evans, carrying with them the 243 pounds of samples they had collected at Taurus-Littrow during traverses that covered more than twenty-one miles. The LM was jettisoned, and this time the impact occurred just west of the landing site, some 475 miles east of the Apollo 15 ALSEP. Its impact was recorded by all four of the previously deployed passive seismometers and the Seismic Profiling experiment deployed on this mission.

While Cernan and Schmitt were on the surface, Ron Evans had been con­ducting the experiments assigned to him. He had also undertaken some experi­ments beginning with the translunar coast phase and would continue making measurements during the return home, almost until reentry. The high activity period, however, was while he was in lunar orbit, where he conducted a suite of experiments similar to those of the previous two missions. Once Cernan and Schmitt were on board and they were on their way back to Earth, Evans carried out an in-flight EVA when he retrieved the film canisters and other data from the SIM bay. Splashdown and recovery were uneventful.

Project Apollo was now a new chapter in the history books. Even with a few glitches, the flight of Apollo 17 had been the most successful mission from a scientific viewpoint. An enormous treasure trove of lunar samples was in the vaults at the Lunar Receiving Laboratory awaiting study. The seismic and laser corner reflector networks were already returning exciting information, as were many other ALSEP experiments. ALSEP central stations were performing up to or beyond their original design goals. But political and public apathy had set in long before the launch of Apollo 17, and the scientific results alone couldn’t convince the decision makers to add more missions. Those of us still working in the Apollo Program Office faced the dismal task of mopping up and closing down an unequaled undertaking. Many of my coworkers had already begun to drift away to other NASA offices or to new work in or outside the government.

Our dreams for lunar exploration never went away; we always hoped that Congress and the Nixon administration would see the error of their ways and provide the funding to reinstate our post-Apollo plans. But in spite of the surprising discoveries made by Apollo 11 and the missions that followed, no national commitment was forthcoming, and the Apollo hardware remaining after Apollo 17 was never used for its original purpose. Some was used for Skylab, some for Apollo-Soyuz, and the other items are lying ignominiously on the ground at museums, like tethered Gullivers, as reminders to millions of visitors each year of Apollo’s magnitude—and perhaps, to some, of oppor­tunities lost. When will man again set foot on the Moon? Or will we bypass the Moon and go directly to Mars? Or will we stay earthbound or in near-Earth orbit for generations?

One afternoon, walking between my office at L’Enfant Plaza and the NASA of­fices at 600 Independence Avenue, I met some of my former Advanced Manned Missions and Apollo colleagues who had recently been assigned to potential future manned space flight programs. We talked briefly about the uncertainties surrounding these programs (none had been officially blessed as the successor to Apollo except for the short-term Skylab and Apollo-Soyuz programs) and discussed where I might find a new job. I was struck by their lack of enthusiasm and their pessimism about their new work as we discussed NASA’s future. It seemed as if almost overnight this marvelous can-do agency had grown old and lost its way. Gray heads were beginning to predominate in all the offices. Instead of looking toward an exciting future, everyone seemed to be scrambling to hang on and find a place to roost. I decided at that moment that it was time to leave NASA and find a new program I could devote my energies to.

Following the lead of Ed Davin and Dick Green, I sent an application to the National Science Foundation and was hired immediately by a new organization called Research Applied to National Needs, which was undertaking research on a wide spectrum of new technologies. Thus began a new career, but never again would I experience the excitement and the sense of achievement that came with being a small part of Project Apollo.